1,282
Views
6
CrossRef citations to date
0
Altmetric
Research Article

Necessary and sufficient conditions for the linearisability of two-input systems by a two-dimensional endogenous dynamic feedback

, & ORCID Icon
Pages 800-821 | Received 08 Jul 2021, Accepted 02 Dec 2021, Published online: 23 Dec 2021

Abstract

We propose easily verifiable necessary and sufficient conditions for the linearisability of two-input systems by an endogenous dynamic feedback with a dimension of at most two.

1. Introduction

The concept of flatness has been introduced in control theory by Fliess, Lévine, Martin and Rouchon, see, e.g. Fliess et al. (Citation1992) and Fliess et al. (Citation1995). For flat systems, many feed-forward and feedback problems can be solved systematically and elegantly, see, e.g. Fliess et al. (Citation1995). Roughly speaking, a nonlinear control system of the form x˙=f(x,u)with dim(x)=n states and dim(u)=m inputs is flat, if there exist m differentially independent functions yj=φj(x,u,u1,,uq), where uk denotes the kth time derivative of u, such that x and u can locally be parameterised by y and its time derivatives. For this flat parameterisation, we write x=Fx(y,y1,,yr1),u=Fu(y,y1,,yr)and refer to it as the parameterising map with respect to the flat output y. If the parameterising map is invertible, i.e. y and all its time derivatives which explicitly occur in the parameterising map can be expressed solely as functions of x and u, the system is exactly linearisable by static feedback. In this case, we call y a linearising output of the static feedback linearisable system. The static feedback linearisation problem has been solved completely, see Jakubczyk and Respondek (Citation1980) and Nijmeijer and van der Schaft (Citation1990). However, for flatness there do not exist easily verifiable necessary and sufficient conditions, except for certain classes of systems, including two-input driftless systems, see Martin and Rouchon (Citation1994) and systems which are linearisable by a onefold prolongation of a suitably chosen control, see Nicolau and Respondek (Citation2017). Necessary and sufficient conditions for (x,u)-flatness of control affine systems with two inputs and four states can be found in Pomet (Citation1997).

It is well known that every flat system can be rendered static feedback linearisable by an endogenous dynamic feedback, and conversely, every system linearisable by an endogenous dynamic feedback is flat. If a flat output is known, such a linearising feedback can be constructed systematically, see, e.g. Fliess et al. (Citation1999). In this contribution, we propose easily verifiable necessary and sufficient conditions for the linearisability of two-input systems by an endogenous dynamic feedback with a dimension of at most two. In Gstöttner et al. (Citation2021b), a sequential test for checking whether a two-input system is linearisable by an endogenous dynamic feedback with a dimension of at most two has been proposed recently. The main idea of the sequential test in Gstöttner et al. (Citation2021b) is to successively split off or add endogenous dynamic feedbacks to the system in such a way that eventually a static feedback linearisable system is obtained and it is shown that the proposed algorithm succeeds if and only if the original system is indeed linearisable by an at most two-dimensional endogenous dynamic feedback. However, a major drawback of this sequential test is that it requires straightening out involutive distributions, which from a computational point of view is unfavourable. The necessary and sufficient conditions which we propose in the present contribution overcome this computational drawback. Instead of a sequence of systems a certain sequence of distributions is constructed, based on which it can be decided whether the two-input system is linearisable by an endogenous dynamic feedback with a dimension of at most two or not. Constructing these distributions and verifying the proposed conditions require differentiation and algebraic operations only.

It turns out that systems which are linearisable by an endogenous dynamic feedback with a dimension of at most two are actually linearisable by a special kind of endogenous dynamic feedback, namely prolongations of a suitably chosen input after a suitable static feedback transformation has been applied to the system. A complete solution for the flatness problem for the class of two-input systems which are linearisable by a onefold prolongation of a suitably chosen control is provided in Nicolau and Respondek (Citation2016b). Two-input systems which are linearisable by a twofold prolongation of a suitable chosen control are considered in Nicolau and Respondek (Citation2016a). However, no complete solution for the flatness problem of this class of systems is provided in Nicolau and Respondek (Citation2016a) due to Assumption 2 therein. In Section 6, we apply our results to some examples, to none of which the results in Nicolau and Respondek (Citation2016a) are applicable. Normal forms for systems which are linearisable by a onefold prolongation can be found in Nicolau and Respondek (Citation2019), and normal forms for control affine two-input systems linearisable by a twofold prolongation have recently been proposed in Nicolau and Respondek (Citation2020). The present contribution is greatly influenced by all these results. The novelty of our contribution is easily verifiable necessary and sufficient conditions for linearisability by a twofold prolongation, covering also the cases to which the results in Nicolau and Respondek (Citation2016a) do not apply. The necessary and sufficient conditions which we propose in this contribution are also a major improvement over the in principal verifiable but computationally inefficient necessary and sufficient conditions in Gstöttner et al. (Citation2021b).

This paper is organised as follows. In Section 2, we introduce the notation used throughout the paper. In Section 3, some preliminaries regarding flatness of two-input systems are presented, and in Section 4, the sequential test from Gstöttner et al. (Citation2021b) is recapitulated briefly. The main results of this contribution are presented in Section 5, and in Section 6, they are applied to practical and academic examples.

2. Notation

Let X be an n-dimensional smooth manifold, equipped with local coordinates xi, i=1,,n. Its tangent bundle is denoted by (T(X),τX,X), for which we have the induced local coordinates (xi,x˙i) with respect to the basis {xi}. We make use of the Einstein summation convention. By xh, we denote the m×n Jacobian matrix of h=(h1,,hm) with respect to x=(x1,,xn). The k-fold Lie derivative of a function φ along a vector field v is denoted by Lvkφ. Let v and w be two vector fields. Their Lie bracket is denoted by [v,w], for the repeated application of the Lie bracket, we use the common notation advkw=[v,advk1w], k1 and adv0w=w. Let D1 and D2 be two distributions. By [v,D1], we denote the distribution spanned by the Lie bracket of v with all basis vector fields of D1, and by [D1,D2] the distribution spanned by the Lie brackets of all possible pairs of basis vector fields of D1 and D2. The first derived flag of a distribution D is denoted by D(1) and defined by D(1)=D+[D,D]. The involutive closure of D is denoted by D¯ and is the smallest involutive distribution which contains D. By C(D), we denote the Cauchy characteristic distribution of D. It is spanned by all vector fields cD which satisfy [c,D]D. Cauchy characteristic distributions are always involutive. The symbols ⊂ and are used in the sense that they also include equality. An integer beneath the symbol ⊂ denotes the difference of the dimensions of the distributions involved, e.g. D1kD2 means that D1D2 and dim(D2)=dim(D1)+k. We make use of multi-indices, in particular by R=(r1,r2) we denote the unique multi-index associated to a flat output of a system with two inputs, where rj denotes the order of the highest derivative of yj needed to parameterise x and u by this flat output, i.e. y[R]=(y1,y11,,yr11,y2,y12,,yr22). Furthermore, we define R±c=(r1±c,r2±c) with an integer c, and #R=r1+r2.

3. Preliminaries

In this section, we summarise some results regarding flatness of two-input systems. Throughout, all functions and vector fields are assumed to be smooth and all distributions are assumed to have locally constant dimension, we consider generic points only. Consider a nonlinear two-input system of the form (1) x˙i=fi(x,u),i=1,,n(1) with dim(x)=n, dim(u)=2 and rank(uf)=2.

Definition 3.1

The two-input system (Equation1) is called flat if there exist two differentially independent functions yj=φj(x,u,u1,,uq), j = 1, 2 and smooth functions Fxi and Fuj such that locally (2) xi=Fxi(y,y1,,yr1), i=1,,nuj=Fuj(y,y1,,yr), j=1,2.(2) The functions yj=φj(x,u,u1,,uq), j = 1, 2 are called the components of the flat output y=φ(x,u,u1,,uq).

Let y=φ(x,u,u1,,uq) be a flat output of (Equation1). We define the multi-index R=(r1,r2) where rj is the order of the highest derivative of the component yj of the flat output which explicitly occurs in (Equation2). This multi-index can be shown to be unique and with this multi-index, the flat parameterisation can be written in the form (3) x=Fx(y[R1])u=Fu(y[R]).(3) The flat parameterisation is a submersion (it degenerates to a diffeomorphism if and only if y is a linearising output). The difference of the dimensions of the domain and the codomain of (Equation3) is denoted by d, i.e. d=#R+2(n+2)=#Rn. In Nicolau and Respondek (Citation2016b) and Nicolau and Respondek (Citation2016a), the number #R+2 is called the differential weight of the flat output. (The differential weight of a flat output with difference d is thus given by n + 2 + d.) The difference d is the minimal dimension of an endogenous dynamic feedback needed to render (Equation1) static feedback linearisable such that y forms a linearising output of the closed-loop system. Such a linearising endogenous feedback can be constructed systematically, see, e.g. Fliess et al. (Citation1999). If we have d = 0, the map (Equation3) degenerates to a diffeomorphism and the system is static feedback linearisable with y being a linearising output. A flat output y is called a minimal flat output if its difference is minimal compared to all other possible flat outputs of the system. We define the difference d of a flat system to be the difference of a minimal flat output of the system. The difference d of a system therefore measures its distance from static feedback linearisability, i.e. d is the minimal possible dimension of an endogenous dynamic feedback needed to render the system static feedback linearisable.

For (Equation1), we define the distributions D0=span{u1,u2} and Di=Di1+[f,Di1], i1 on the state and input manifold X×U, where f=fi(x,u)xi.

Theorem 3.2

The two-input system (Equation1) is linearisable by static feedback if and only if all the distributions Di are involutive and dim(Dn1)=n+2.

For a proof of this theorem, we refer to Nijmeijer and van der Schaft (Citation1990). For a system which meets the conditions of Theorem 3.2, the linearising outputs can be computed as follows. Let s be the smallest integer such that Ds=T(X×U). In case of dim(Ds1)=n (i.e. Ds1 is of codimension 2), the sequence of involutive distributions is of the form D02D122Ds12T(X×U)and linearising outputs are all pairs of functions (φ1,φ2) which satisfy span{dφ1,dφ2}=Ds1. However, if dim(Ds1)=n+1 (i.e. Ds1 is of codimension 1), the sequence is of the form D02D122Dl12Dl1Dl+111Ds11T(X×U),i.e. there exists an integer l from which on the sequence grows in steps of one. Linearising outputs are then all pairs of functions (φ1,φ2) which satisfy span{dφ1}=Ds1 and span{dφ1,dLfφ1,,dLfslφ1,dφ2}=Dl1.

The sequential test for flatness with d2 proposed in Gstöttner et al. (Citation2021b), as well as the distribution test for flatness with d2 which we propose in this contribution, both rely on the following crucial result regarding flat two-input systems with d2.

Theorem 3.3

A system (Equation1) with d2 can be rendered static feedback linearisable by d-fold prolonging a suitably chosen (new) input after a suitable input transformation u¯=Φu(x,u) has been applied.

A proof of this result is provided in Appendix 1.

4. Sequential test

In this section, we briefly recapitulate the main idea of the necessary and sufficient condition for flatness with d2 in form of the sequential test proposed in Gstöttner et al. (Citation2021b). For details, proofs and examples, we refer to Gstöttner et al. (Citation2021b). Let y be a minimal flat output with difference 0<d2 of the system (Equation1). It can be shown that the assumption 0<d2 implies the existence of an input transformation u¯=Φu(x,u) such that the flat parameterisation of the new inputs by the flat output y is of the form u¯1=F¯u1(y[R1]), u¯2=F¯u2(y[R]) (where F¯u=Φu(Fx,Fu)). Consider the system obtained by onefold prolonging u¯1, i.e.  x˙=f(x,Φˆu(x,u¯))=f¯(x,u¯)u¯˙1=u¯11,with the state (x,u¯1) and the input (u¯11,u¯2), and where u=Φˆ(x,u¯) is the inverse of the transformation u¯=Φ(x,u). The flat output y of the original system is also a flat output of the prolonged system (and conversely, it can be shown that every flat output of the prolonged system is also a flat output of the original system). Since u¯1=F¯u1(y[R1]), we have u¯11=F¯u11(y[R]) and thus, the domain of the parameterising map of the prolonged system with respect to the flat output y is still of dimension #R+2, but its codomain grew by one, i.e. y as a minimal flat output of the prolonged system has a difference of d−1 only. The main idea of the sequential test in Gstöttner et al. (Citation2021b) is to find such an input (they can indeed be found systematically), prolong it in order to obtain a system whose difference is d−1 (where d2 is the difference of the original system), and since by assumption d2, after at most two such steps the procedure must yield a static feedback linearisable system. Otherwise, the original system must have had a difference of d3.

When applying the sequential test to a system (Equation1), in every step a new system is derived by either splitting off a two-dimensional endogenous dynamic feedback or by adding a one-dimensional endogenous dynamic feedback (in form of a one-fold prolongation of a certain input). How the next system is derived from the current one is decided based on the distributions D0=span{u1,u2} and D1=D0+[f,D0] of the current system. If D1 is involutive, it can be straightened out by a suitable state transformation x¯=Φx(x) in order to obtain a decomposition of the system into the form (4) Σ2:x¯˙2i2=f¯2i2(x¯2,x¯1), i2=1,,n2Σ1:x¯˙1i1=f¯1i1(x¯2,x¯1,u), i1=1,2.(4) The procedure is then continued with the subsystem Σ2 with the state x¯2 and the input x¯1, i.e. we split off a two-dimensional endogenous dynamic feedback. It follows that Σ2 has the same flat outputs with the same differences as the original system.

If D1 is non-involutive but D0C(D1), it can be shown that the system allows an affine input representation (AI representation) x˙=a(x)+b1(x)u1+b2(x)u2with a non-involutive input distribution span{b1,b2}. Based on such an AI representation, an input transformation u¯j=mlj(x)ul, j, l = 1, 2 can be derived such that if the system indeed has a difference of d2, the system obtained by onefold prolonging the new input u¯1 has a difference of d−1, i.e. in such a step a one-dimensional endogenous dynamic feedback is added to the system, and under the assumption d2, it can be shown that the feedback modified system has a difference of d−1 only.

Finally, if D1 is non-involutive and D0C(D1), the system allows at most a so-called partial affine input representation (PAI representation) x˙=a(x,u¯1)+b(x,u¯1)u¯2.This form was introduced in Schlacher and Schöberl (Citation2013). In Kolar et al. (Citation2016), it has been shown that the existence of a PAI representation is a necessary condition for flatness. For two input systems, an input transformation such that the system takes PAI form can be derived systematically (provided the system indeed allows a PAI representation, otherwise, we can conclude that the system is not flat). It can be shown that a system which does not allow an AI representation allows at most two fundamentally different PAI representations and that in case of d2, the non-affine occurring inputs u¯1 of these possibly existing two PAI representations are the candidates for inputs whose flat parameterisation with respect to a minimal flat output involves derivatives up to order R−1 only. So in this case, the procedure is continued with the system obtained by onefold prolonging the non-affine occurring input u¯1 (if the system indeed allows two fundamentally different PAI representations, we have to continue the procedure with both of them, i.e. there may occur a branching point).

As already mentioned, the sequential test has the drawback that it requires straightening out involutive distributions to achieve decompositions of the form (Equation4). In fact, also the explicit computation of an input transformation such that a system takes PAI form requires straightening out an involutive distribution.

5. Main results

In this section, we present our main results, which are easily verifiable necessary and sufficient conditions for flatness with a difference of d2 in the form of Theorem 5.1 for the case d = 2 and Theorem 5.2 for the case d = 1 below. These necessary and sufficient conditions overcome the computational drawbacks of the sequential test described in the previous section. Instead of a sequence of systems, a certain sequence of distributions is constructed. The distributions constructed when applying these theorems are actually closely related to the distributions D0 and D1 of the individual systems constructed in the sequential test on basis of which in the sequential test it is decided how the next system is computed from the current one. There is actually a one-to-one correspondence between the sequential test and the conditions of Theorems 5.1 and 5.2. One could prove these theorems via this one-to-one correspondence, however, in this contribution we provide self-contained proofs which do not rely on the sequential test. A detailed proof of Theorem 5.1 is provided in Section 7, and for Theorem 5.2, a brief sketch of a proof is provided in Section 8. As already mentioned, the case d = 1 has been solved completely in Nicolau and Respondek (Citation2016b). Below, we explain how our necessary and sufficient conditions for the case d = 1 in the form of Theorem 5.2 are related with those provided in Nicolau and Respondek (Citation2016b). The computation of flat outputs with d2 of systems which meet our conditions for flatness with d2 is addressed in Section 5.2.

Assume that the system (Equation1) is not static feedback linearisable. We then have the involutive distribution D0=span{u1,u2} and can calculate the distributions Di=Di1+[f,Di1], i=1,,k1 where k1 is defined to be the smallest integer such that Dk1 is non-involutive (its existence is assured by the assumption that the system is not static feedback linearisable).

Theorem 5.1

The system (Equation1) is flat with a difference of d = 2 if and only if:

  1. The distributions Di, i=1,,k1 have the dimensions dim(Di)=2(i+1).

  2. Either Dk11C(Dk1) and then:

    1. Either dim(D¯k1)=dim(Dk1)+1 and dim([f,Dk1]+D¯k1)=dim(D¯k1)+1. Define Ek1+1=D¯k1 and continue with item 3b.

    2. Or dim(D¯k1)=dim(Dk1)+2 and [f,C(Dk1(1))]Dk1(1). Define Ek1+1=Dk1(1) and continue with item 4a.II with k2=k1+1.

  3. Or Dk11C(Dk1) and then there exists a vector field vcDk11, vcDk12 (take Dk12=0 if k1=1) such that the distributions Ek11=Dk12+span{vc} and Ek1=Dk11+span{[vc,f]} meet Ek11C(Ek1).Footnote1

  4. Either Ek1 is non-involutive (only 2b. can yield a non-involutive distribution Ek1) and then

    1. dim(E¯k1)=dim(Ek1)+1.

    2. dim([f,Ek1]+E¯k1)=dim(E¯k1)+1. Define Fk1+1=E¯k1 and continue with item 5.

  5. Or Ek1 is involutive and then

    1. There exists a minimal integer k2 such that Ek2 is non-involutive, where Ei=Ei1+[f,Ei1].

    2. The distributions Ei have the dimensions dim(Ei)=2i+1 for i=k1+1,,k2.

  6. Either Ek21C(Ek2) and then

    1. dim(E¯k2)=dim(Ek2)+1.

    2. Either E¯k2=T(X×U), or dim([f,Ek2]+E¯k2)=dim(E¯k2)+1. Define Fk2+1=E¯k2.

  7. Or Ek21C(Ek2) and then Fk2=Ek21+C(Ek2) is involutive.

  8. All the distributions Fi=Fi1+[f,Fi1] are involutive and there exists an integer s such that Fs=T(X×U).

In Theorem 5.1, we have several junctions, which is graphically illustrated in Figure .

Figure 1. Overview of the possible paths in Theorem 5.1.

Figure 1. Overview of the possible paths in Theorem 5.1.

Regarding flatness with a difference of d = 1, we have the following result.

Theorem 5.2

The system (Equation1) is flat with a difference of d = 1 if and only if:

  1. The distributions Di, i=1,,k1 have the dimensions dim(Di)=2(i+1).

  2. Either Dk11C(Dk1) and then

    1. dim(D¯k1)=dim(Dk1)+1.

    2. Either D¯k1=T(X×U), or dim([f,Dk1]+D¯k1)=dim(D¯k1)+1. Define Ek1+1=D¯k1.

  3. Or Dk11C(Dk1) and then Ek1=Dk11+C(Dk1) is involutive.

  4. All the distributions Ei=Ei1+[f,Ei1] are involutive and there exists an integer s such that Es=T(X×U).

In Theorem 5.2, we have exactly one junction, which is graphically illustrated in Figure .

Figure 2. Overview of the possible paths in Theorem 5.2.

Figure 2. Overview of the possible paths in Theorem 5.2.

Note that the items 1. to 3. of Theorem 5.2 in fact coincide with the items 3b. to 5. of Theorem 5.1 when Di and k1 are replaced by Ei and k2. In Nicolau and Respondek (Citation2016b), necessary and sufficient conditions for the case d = 1 are provided via Theorems 3.3 and 3.4 therein. These theorems are stated for the control affine case, but this is actually no restriction. It can be shown that the control affine system obtained by prolonging both inputs of a general nonlinear control system of the form (Equation1), i.e.  x˙=f(x,u)u˙1=u11u˙2=u12with the state (x,u1,u2) and the input (u11,u12) is flat with a certain difference d if and only if the original system is flat with the same difference d. In fact, a flat output of the original system with a certain difference d is also a flat output of the prolonged system with the same difference d and vice versa. In Theorem 5.2, we have exactly one junction (see Figure ). This distinction of cases between Dk11C(Dk1) and Dk11C(Dk1) is motivated by the sequential test of the previous section (it corresponds to the distinction between AI form and PAI form in the sequential test). Our necessary and sufficient conditions for flatness with d = 1 as stated in Theorem 5.2 are very similar to those stated in Nicolau and Respondek (Citation2016b). The main difference is in fact that in Nicolau and Respondek (Citation2016b) a distinction of cases is made between D¯k1T(X×U) (Theorem 3.3 therein) and D¯k1=T(X×U) (Theorem 3.4 therein), instead of a distinction of cases between Dk11C(Dk1) and Dk11C(Dk1) as it is done in Theorem 5.2.

It can easily be shown that all the distributions and all the conditions in Theorems 5.1 and 5.2 are invariant with respect to regular input transformations u¯=Φu(x,u). Although the vector field f=fi(x,u)xi associated with (Equation1) and the vector field f¯=f¯i(x,u¯)xi associated with the feedback modified system x˙i=f¯i(x,u¯), where f¯i(x,u¯)=fi(x,Φˆu(x,u¯)) with the inverse u=Φˆu(x,u¯) of the input transformation u¯=Φu(x,u), are in general only equal modulo D0=span{u}=span{u¯}, the distributions Di, Ei and Fi constructed from them in the above theorems coincide.

5.1. Verification of the conditions

All the conditions of Theorems 5.1 and 5.2 are easily verifiable and require differentiation and algebraic operations only. Item 2b. of Theorem 5.1 can be verified as follows. Choose any pair of vector fields v1,v2Dk11 such that Dk11=Dk12+span{v1,v2}. Any vector field vcDk11, vcDk12 can then be written as a non-trivial linear combination vc=α1v1+α2v2 mod Dk12. Since Ek11=Dk12+span{vc} and Ek1=Dk11+span{[vc,f]}, the condition Ek11C(Ek1) implies vcC(Ek1) and in turn [vc,[vc,f]]Ek1. Since for any vc by construction we have Ek1Dk1, the condition [vc,[vc,f]]Ek1 implies [vc,[vc,f]]Dk1, which yields the necessary condition (5) (α1)2[v1,[v1,f]]+2α1α2[v1,[v2,f]]+(α2)2[v2,[v2,f]]!Dk1(5) (where we have used that Dk12C(Dk1), which can be shown based on the Jacobi identity). The following lemma states a crucial property of (Equation5).

Lemma 5.1

The condition (Equation5) admits at most two independent non-trivial solutions αj(x,u).

A proof of this lemma is provided in Appendix 1. (A similar result has been proven in Gstöttner et al. (Citation2020b) in the context of a certain structurally flat triangular form. The result which we prove here is more general.) Since (Equation5) admits at most two independent non-trivial solutions, there also exist at most two vector fields vc=α1v1+α2v2 which are not collinear modulo Dk12 and meet the above criterion. Since in Ek11=Dk12+span{vc} and Ek1=Dk11+span{[vc,f]} only the direction of vc modulo Dk12 matters, there also exist at most two distinct such pairs of distributions.

5.2. Computation of flat outputs

Theorems 5.1 and 5.2 allow us to check whether a system is flat with a difference of d2. Regarding the computation of the corresponding flat outputs with d2 we have the following result.

Theorem 5.3

Assume that the system (Equation1) meets the conditions of Theorems 5.1 or 5.2. Flat outputs with d = 1 or d = 2 can then be determined from the sequence of involutive distributions Ei or Fi the same way as linearising outputs are determined from the sequence of involutive distributions Di involved in the test for static feedback linearisability in Theorem 3.2.

Proof.

The sufficiency parts of the proofs of Theorems 5.1 and 5.2 are done constructively. Based on the distributions involved in the conditions of these theorems, for each case a certain coordinate transformation such that the system takes a structurally flat triangular form is derived. For details, see the sufficiency parts of the proofs of Theorems 5.1 and 5.2.

Remark 5.1

Above we have defined the difference d of a system as the difference of a minimal flat output and thus the minimal possible dimension of an endogenous dynamic feedback needed to render the system static feedback linearisable. Although a system with a difference of d = 1 may also be linearised by an endogenous dynamic feedback of a higher dimension, in particular a two-dimensional one, it is important to note that for such a system only the conditions of Theorem 5.2 are met but those of Theorem 5.1 are not met. As a consequence, the computation of flat outputs as described in Theorem 5.3 only yields minimal flat outputs.

It may happen that for the computation of flat outputs as stated in Theorem 5.3, a distribution which does not explicitly occur in Theorems 5.1 or 5.2 is needed. To be precise, if in Theorem 5.2 the conditions 2a apply and we have Ek1+1=T(X×U) or Ek1+11Ek1+2, we have to construct an involutive distribution Ek1 which satisfies Dk111Ek11Dk1 and [f,Ek1]D¯k1 for the computation of flat outputs. Such a distribution always exists and except for the case Ek1+1=T(X×U), it is also unique. The construction is as follows. In case of Ek1+1=T(X×U), choose any function ψ whose differential dψ0 annihilates Dk11 and choose furthermore any pair of vector fields v1,v2 which complete Dk11 to Dk1, i.e. such that Dk1=Dk11+span{v1,v2}. The distribution Ek1=Dk11+span{(dψv2)v1(dψv1)v2} can then be shown to be involutive. Different choices for ψ lead in general to different distributions Ek1, but different choices for v1,v2 have no effect. (A flat output with d = 1 is then formed by any pair of functions (φ1,φ2) satisfying span{dφ1,dφ2}=Ek1, where we can always choose one of the components equal to ψ since by construction dψEk1.)

In case of Ek1+1T(X×U), the condition dim([f,Dk1]+D¯k1)=dim(D¯k1)+1 implies the existence of a vector field vDk1, vDk11 such that [v,f]D¯k1. The direction of v is unique modulo Dk11, and it follows that the distribution Ek1=Dk11+span{v} constructed from it is involutive. (A flat output with d = 1 is then formed by any pair of functions (φ1,φ2) satisfying span{dφ1}=Es1 and span{dφ1,dLfφ1,,dLfsk11φ1,dφ2}=Ek1.)

Similarly, if in Theorem 5.1 the conditions 4a (or 2a.B and 4a.II, in which case we define Ek21=C(Dk1(1))) apply and we have Fk2+1=T(X×U) or Fk2+11Fk2+2, we have to construct an involutive distribution Fk2 which satisfies Ek211Fk21Ek2 and [f,Fk2]E¯k2 for the computation of flat outputs. This can be done as just explained for Theorem 5.2, simply replace Dk11 and Dk1 by Ek21 and Ek2.

Finally, if in Theorem 5.1 the conditions 3a apply and we have Fk1+11Fk1+2, we have to construct an involutive distribution Fk1 which satisfies Ek111Fk11Ek1 and [f,Fk1]E¯k1 for the computation of flat outputs. The construction is again the same as just explained for Theorem 5.2, simply replace Dk11 and Dk1 by Ek11 and Ek1. Since we always have E¯k1T(X×U) in this case, the distribution Fk1 is unique.

6. Examples

In the following, we apply our results to several examples. Most of these examples are well-known benchmark examples for which flat outputs have been found on the basis of physical considerations (see, e.g. Martin et al. (Citation1996)) or constructive procedures (see, e.g. Schöberl and Schlacher (Citation2014)). Our results allow for a systematic computation of flat outputs for these examples. We focus on the case d = 2, as the novelty of this contribution are the results regarding the case d = 2, which also cover the cases which cannot be handled with the results in Nicolau and Respondek (Citation2016a) (none of the following examples meets Assumption 2 therein). It should again be pointed out that from a computational point of view, the necessary and sufficient conditions for flatness with d2 in the form of Theorems 5.1 and 5.2 are a major improvement over the necessary and sufficient conditions in form of the sequential test proposed in Gstöttner et al. (Citation2021b).

6.1. VTOL Aircraft

Consider the model of a planar VTOL aircraft (6) x˙=vxz˙=vzθ˙=ωv˙x=ϵcos(θ)u2sin(θ)u1v˙z=cos(θ)u1+ϵsin(θ)u21ω˙=u2.(6) This system is also treated in e.g. Fliess et al. (Citation1999), Gstöttner et al. (Citation2020b), Gstöttner et al. (Citation2021b), Schöberl et al. (Citation2010) or Schöberl and Schlacher (Citation2011). The distributions D0=span{u1,u2} and D1=span{u1,u2,sin(θ)vx+cos(θ)vz,ϵcos(θ)vx+ϵsin(θ)vz+ω}are involutive, but D2=span{u1,u2,sin(θ)vx+cos(θ)vz,ϵcos(θ)vx+ϵsin(θ)vz+ω,sin(θ)xcos(θ)zωcos(θ)vxωsin(θ)vz,ϵcos(θ)x+ϵsin(θ)z+θ+ϵωsin(θ)vxϵωcos(θ)vz}is non-involutive, so we have k1=2. With these distributions, item 1 of Theorem 5.1 is met. The Cauchy characteristic distribution of D2 follows as C(D2)=span{u1,u2}=D0 and thus D1C(D2). So we are in item 2b and have to construct a vector field vcD1, vcD0 such that E1C(E2) with E1=D0+span{vc} and E2=D1+span{[vc,f]}. So we set vc=α1v1+α2v2 with arbitrary vector fields v1,v2 which complete D0 to D1, e.g. v1=sin(θ)vx+cos(θ)vz, v2=ϵcos(θ)vx+ϵsin(θ)vz+ω, and determine α1,α2 from (Equation5), which in this particular case reads (7) 2α1α2(cos(θ)vx+sin(θ)vz)+(α2)2(2ϵsin(θ)vx2ϵcos(θ)vz)!D2,(7) and admits the two independent solutions α1=λ, α2=0 and α1=0, α2=λ, both with an arbitrary function λ0. We can choose λ=1 since only the direction of vc=α1v1+α2v2 matters, so we simply have vc,1=v1=sin(θ)vx+cos(θ)vz and vc,2=v2=ϵcos(θ)vx+ϵsin(θ)vz+ω. With both of these vector fields we indeed have E1,jC(E2,j), where E1,j=D0+span{vc,j} and (8) E2,1=span{u1,u2,sin(θ)vx+cos(θ)vz,ϵcos(θ)vx+ϵsin(θ)vz+ω,sin(θ)xcos(θ)zωcos(θ)vxωsin(θ)vz}(8) and (9) E2,2=span{u1,u2,sin(θ)vx+cos(θ)vz,ϵcos(θ)vx+ϵsin(θ)vz+ω,ϵcos(θ)x+ϵsin(θ)z+θ+ϵωsin(θ)vxϵωcos(θ)vz}.(9) We thus have a branching point and have to continue with both of these distributions (we will be able to discard one of them in just a moment). The distributions E2,j are non-involutive, so we are in item 3a. Both meet the condition 3a.I, i.e. dim(E¯2,j)=dim(E2,j)+1. However, 3a.II is only satisfied by E2,2, for E2,1 we have dim([f,E2,1]+E¯2,1)=dim(E¯2,1)+2, so we can discard this branch and continue with E2,2 only, where for ease of notation we drop the second subscript from now on, i.e. E2=E2,2. According to item 3a.II, we have F3=E¯2=span{u1,u2,ω,vx,vz,ϵcos(θ)x+ϵsin(θ)z+θ}. Continuing this sequence as stated in item 5, we obtain F4=T(X×U), so the conditions of item 5 are also met and we conclude that the system (Equation6) is flat with a difference of d = 2. In conclusion, the system meets the items 1, 2b, 3a, 5 (which corresponds to the 4th path from the left in Figure ).

According to Theorem 5.3, flat outputs with d = 2 of the VTOL aircraft can thus be computed from the distributions Fi the same way as linearising outputs are determined from the sequence of involutive distributions involved in the test for static feedback linearisability. We have F32F4=T(X×U), flat outputs with d = 2 are thus all pairs of functions (φ1,φ2) satisfying span{dφ1,dφ2}=F3. We have F3=span{dxϵcos(θ)dθ,dzϵsin(θ)dθ} and thus e.g. φ1=xϵsin(θ), φ2=z+ϵcosθ.

The sufficiency part of the proof of Theorem 5.1 is done constructively. Based on the distributions involved in the conditions of the theorem, a coordinate transformation such that the system takes a structurally flat triangular form is derived. To aid understanding the sufficiency part of the proof of Theorem 5.1, we explicitly derive this coordinate transformation for this particular example. We have the following sequence of involutive distributions D0E1F2F3F4=T(X×U),where E1=D0+span{ϵcos(θ)vx+ϵsin(θ)vz+ω} and F2=E1+span{ϵcos(θ)x+ϵsin(θ)z+θϵωsin(θ)vx+ϵωcos(θ)vz} is the unique involutive distribution determined by the conditions E11F21E2 and [f,F2]E¯2. In order to straighten out all these distributions simultaneously, we apply the state transformation x¯42=xϵsin(θ)x¯41=z+ϵcos(θ)x¯32=vxϵωcos(θ)x¯31=vzϵωsin(θ)x¯21=θx¯11=ω.In these coordinates, we have x¯˙42=x¯32x¯˙41=x¯31x¯˙32=sin(x¯21)(ϵ(x¯11)2u1)x¯˙31=cos(x¯21)(u1ϵ(x¯11)2)1x¯˙21=x¯11x¯˙11=u2,which is of the form (Equation26). Normalising the third equation by introducing u¯2=sin(x¯21)(ϵ(x¯11)2u1), u¯1=u2, we obtain x¯˙42=x¯32x¯˙41=x¯31x¯˙32=u¯2x¯˙31=cot(x¯21)u¯21x¯˙21=x¯11x¯˙11=u¯1,which is of the form (Equation25). The top variables x¯42, x¯41 form a flat output of (Equation6) with a difference of d = 2. Similar triangular forms for the VTOL aircraft have been derived in Gstöttner et al. (Citation2020b) and in Nicolau and Respondek (Citation2020).

6.2. Academic Example 1

Consider the system (10) x˙1=u1x˙2=u2x˙3=sin(u1u2),(10) also considered in Gstöttner et al. (Citation2020b), Gstöttner et al. (Citation2021b), Lévine (Citation2009) and Schöberl and Schlacher (Citation2014). The distribution D0=span{u1,u2} is involutive, D1=span{u1,u2,x1+1u2cos(u1u2)x2,x2u1(u2)2cos(u1u2)x3}is non-involutive, so we have k1=1 and item 1 is met. We have C(D1)=0 and thus D0C(D1), so we are in item 2b and have to construct a vector field vcD0=span{u1,u2}, i.e. vc=α1u1+α2u2, such that E0C(E1) where E0=span{vc} and E1=D0+span{[vc,f]}. For that, we solve (Equation5), which in the particular case yields (α1)2sin(u1u2)(u2)2+2α1α2(cos(u1u2)u2sin(u1u2)u1)u2+(α2)2(sin(u1u2)u12cos(u1u2)u2)u1=!0,and admits the two independent solutions α1=λu1, α2=λu2 and α1=λ(u1tan(u1u2)2u2), α2=λu2tan(u1u2), both with an arbitrary function λ0, e.g. λ=1 since only the direction of vc=α1u1+α2u2 matters. It can easily be checked that only the vector field vc=u1u1+u2u2, obtained from the first solution, satisfies vcC(E1), where E1=D0+span{[vc,f]} follows as E1=span{u1,u2,u1x1+u2x2}. The distribution E1 is non-involutive, so we are in item 3a. We have E¯1=span{u1,u2,x1,x2}. The conditions dim(E¯1)=dim(E1)+1 and dim([f,E1]+E¯1)=dim(E¯1)+1 are met. According to item 3a.II, we have F2=E¯1. Continuing this sequence as stated in item 5, we obtain F3=T(X×U), so the conditions of item 5 are also met and we conclude that the system (Equation10) is flat with a difference of d = 2. In conclusion, the system meets the items 1, 2b, 3a, 5 (which corresponds to the 4th path from the left in Figure ).

According to Theorem 5.3, flat outputs with d = 2 of the system (Equation10) can thus be computed from the distributions Fi the same way as linearising outputs are determined from the sequence of involutive distributions involved in the test for static feedback linearisability. However, in contrast to the previous example, not all the distributions needed for computing flat outputs explicitly occur in Theorem 5.1. We have F21F3=T(X×U), from which we only find one component φ1 of the flat outputs, i.e. span{dφ1}=F2=span{dx3}, from which e.g. φ1=x3 follows. In order to find a second component φ2, we have to complete the sequence to F12F21F3=T(X×U), as stated below Theorem 5.3, and then we find φ2 from span{dφ1,dLfφ1,dφ2}=F1. We have E1=E0+span{u1,u1x1+u2x2} and it follows that [f,u1x1+u2x2]E¯k1. We thus have F1=E0+span{u1x1+u2x2}=span{u1u1+u2u2,u1x1+u2x2} and in turn F1=span{dx3,u2du1u1du2,u2dx1u1dx2}, from which a possible second component follows as φ2=x1x2u1u2. In conclusion, we have derived the flat output φ1=x3, φ2=x1x2u1u2.

Consider the following two examples: (11) x˙1=u1x˙2=u2x˙3=u1u2,x˙1=u1x˙2=u2x˙3=u1u2,(11) also treated in, e.g. Schöberl and Schlacher (Citation2010) and Schöberl and Schlacher (Citation2015). The first one of these two systems can be shown to be flat with a difference of d = 2, where again the items 1, 2b, 3a, 5 (which again corresponds to the 4th path from the left in Figure ) are met. Item 2b yields two different pairs of distributions E0 and E1 for this system, namely E0,1=span{u1}, E1,1=span{u1,u2,x1+u2x3} and E0,2=span{u2}, E1,2=span{u1,u2,x2+u1x3}. For both branches, the items 3a and 5 are met, and we obtain (x2,x3x1u2) and (x1,x3x2u1) as possible flat outputs with d = 2.

However, the second system in (Equation11) is an example which does not meet the conditions for flatness with d2.Footnote2 Hence, we can conclude that if the system is flat, it must have a difference of d3. (The system is indeed flat with a difference of d = 3, in e.g. Schöberl and Schlacher (Citation2015) a corresponding flat output with d = 3 has been derived.)

6.3. Academic Example 2

Consider the system (12) x˙1=arcsin(u1+u2x2)x4x˙2=x4x˙3=u1x˙4=u2.(12) The distribution D0=span{u1,u2} is involutive, but D1=span{u1,u2,x1+(x2)2(u1+u2)2x3,x1+(x2)2(u1+u2)2x4}is non-involutive, so we have k1=1 and item 1 is met. We have D0D1, so we are in item 2b and have to construct a vector field vcD0, i.e. vc=α1u1+α2u2, such that E0C(E1) where E0=span{vc} and E1=D0+span{[vc,f]}. For that, we solve (Equation5), which in the particular case yields (α1+α2)2=!0and has the up to a multiplicative factor unique solution α1=1, α2=1. The vector field vc=u1u2 obtained from this solution indeed meets vcC(E1) with E1=D0+span{[vc,f]}=span{u1,u2,x3x4}. The distribution E1 is involutive, so we are in item 3b. The distribution E2=span{u1,u2,x1+(x2)2(u1+u2)2x3,x1+(x2)2(u1+u2)2x4,x1x2}is non-involutive, i.e. k2=2 and item 3b.II is also met. We have E1C(E2), so we are in item 4b. According to this item, we have F2=E1+C(E2), which evaluates to F2=span{u1,u2,x1x2,x3x4}and is indeed involutive. Continuing this sequence as stated in item 5, we obtain F3=T(X×U), so the conditions of item 5 are also met and we conclude that the system (Equation12) is flat with a difference of d = 2. In conclusion, the system meets the items 1, 2b, 3b, 4b, 5 (which corresponds to the 6th path from the left in Figure ).

According to Theorem 5.3, flat outputs with d = 2 of this system are all pairs of functions (φ1,φ2) satisfying span{dφ1,dφ2}=F2. We have F2=span{dx1+dx2,dx3+dx4}, and thus, e.g. φ1=x1+x2 and φ2=x3+x4.

6.4. Coin on a moving table

Consider the following model of a coin rolling on a rotating table: (13) x˙=Ωcos(θ)(xsin(θ)ycos(θ))+Rcos(θ)u2y˙=Ωsin(θ)(xsin(θ)ycos(θ))+Rsin(θ)u2θ˙=u1ϕ˙=u2,(13) taken from Kai (Citation2006) and also considered in, e.g. Li et al. (Citation2013) and Li et al. (Citation2016). The distribution D0=span{u1,u2} is involutive, D1=span{u1,u2,θ,Rcos(θ)x+Rsin(θ)y+ϕ}is non-involutive, so we have k1=1 and item 1 is met. We have C(D1)=D0, so we are in item 2a. We furthermore have dim(D¯1)=dim(D1)+2 (the involutive closure of D1 is given by D¯1=T(X×U)), so we are in the subcase 2a.B. The first derived flag of D1 is given by D1(1)=span{u1,u2,θ,Rcos(θ)x+Rsin(θ)y+ϕ,sin(θ)xcos(θ)y},its Cauchy characteristic distribution follows as C(D1(1))=span{u1,u2,Rcos(θ)x+Rsin(θ)y+ϕ} and the condition [f,C(D1(1))]D1(1) is indeed met. By definition we have E2=D1(1) and have to continue with item 4a.II. The item 4a.II is indeed met with E¯2=T(X×U) and thus, also 5 is met (with F3=E¯2=T(X×U)), proving that the system is flat with a difference of d = 2. In conclusion, the system meets the items 1, 2a.B, 4a.II, 5 (which corresponds to the 3rd path from the left in Figure ).

According to Theorem 5.3 (see also below Theorem 5.3), flat outputs with d = 2 are all pairs of functions (φ1,φ2) satisfying span{dφ1,dφ2}=F2, where F2 is an arbitrary involutive distribution satisfying E11F21E2 with E1=C(D1(1)). To construct such a distribution F2, we first choose an arbitrary function ψ whose differential dψ0 annihilates E1, e.g. ψ=θ, and a pair of vector fields v1,v2 which completes E1 to E2, e.g. v1=θ, v2=sin(θ)xcos(θ)y. A possible distribution F2 is then given by F2=E1+span{(dψv2)v1(dψv1)v2}=E1+span{sin(θ)xcos(θ)y}, which yields φ1=θ, φ2=Rϕxcos(θ)ysin(θ) as a possible flat output.

Remark 6.1

Theorems 5.1 and 5.2 together cover the whole class of two-input systems which are linearisable by an at most two-dimensional endogenous dynamic feedback. On the basis of the above examples, we have demonstrated that the conditions of these theorems are indeed systematically verifiable and require differentiation and algebraic operations only. Of course it is possible to try simpler sufficient conditions for flatness first. For example, for the coin on a moving table example, the same flat output could be obtained by applying the procedure proposed in Fossas and Franch (Citation2000). That is, the system (Equation13) becomes static feedback linearisable by a twofold prolongation of the original input u1. The flat output derived above is a linearising output of this prolonged system.

7. Proof of Theorem 5.1

Necessity

Consider a two-input system of the form (Equation1) and assume that it is flat with a difference of d = 2. Let k1 be defined as in Theorem 5.1, i.e. the smallest integer such that Dk1 is non-involutive (if Dk1 would not exist, the system would be static feedback linearisable, which is a contradiction to the assumption that d = 2). According to Theorem 3.3, there exists an input transformation u¯=Φu(x,u) with inverse u=Φˆ(x,u¯) such that the system obtained by twofold prolonging the new input u¯1, i.e. the system (14) x˙=f(x,Φˆu(x,u¯))=f¯(x,u¯)u¯˙1=u¯11u¯˙11=u¯21(14) with the state (x,u¯1,u¯11) and the input (u¯21,u¯2), is static feedback linearisable. Therefore, according to Theorem 3.2, with the vector field fp=f¯i(x,u¯)xi+u¯11u¯1+u¯21u¯11 the distributions (15) Δ0=span{u¯21,u¯2}Δ1=span{u¯21,u¯11,u¯2,[fp,u¯2]}Δ2=span{u¯21,u¯11,u¯1,u¯2,[fp,u¯2],adfp2u¯2}Δ3=span{u¯21,u¯11,u¯1,u¯2,[fp,u¯1],[fp,u¯2],adfp2u¯2,adfp3u¯2}Δs=span{u¯21,u¯11,u¯1,u¯2,x}(15) are all involutive (that the integer s in the last line of (Equation15) coincides with s in item 5 is shown later). Some simplifications can be made, i.e. there exist simpler bases for these distributions, as the following lemma asserts. The proposed bases reveal useful relations between the distributions Δi and Di.

Lemma 7.1

The distributions (Equation15) can be simplified to either

(16) Δ0=span{u¯21,u¯2}Δ1=span{u¯21,u¯11,u¯2,[f¯,u¯2]}Δ2=span{u¯21,u¯11,u¯1,u¯2D0,[f¯,u¯2],adf¯2u¯2}Δ3=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯1],[f¯,u¯2]D1,adf¯2u¯2,adf¯3u¯2}Δk1=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯k12u¯1,adf¯k12u¯2Dk12,adf¯k11u¯2,adf¯k1u¯2}Δk1+1=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯k11u¯1,adf¯k11u¯2Dk11,adf¯k1u¯2,adf¯k1+1u¯2}Δk1+2=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯k1u¯1,adf¯k1u¯2Dk1,adf¯k1+1u¯2,adf¯k1+2u¯2}Δs=span{u¯21,u¯11,u¯1,u¯2,x}.(16)

or (17) Δ0=span{u¯21,u¯2}Δ1=span{u¯21,u¯11,u¯2,[f¯,u¯2]}Δ2=span{u¯21,u¯11,u¯1,u¯2D0,[f¯,u¯2],[u¯1,[f¯,u¯2]]}Δ3=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯1],[f¯,u¯2]D1,[u¯1,[f¯,u¯2]],[f¯,[u¯1,[f¯,u¯2]]]u¯1,u¯2,[f¯,u¯1],[f¯,u¯2]D1}Δs=span{u¯21,u¯11,u¯1,u¯2,x}.(17)

Remark 7.1

It follows from the proof of Lemma 7.1, which is provided in Appendix 2, that in case of k12 we always have the form (Equation16). The form (Equation17) is only relevant if in the case k1=1 we have adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}. In all other cases, the distributions Δi are indeed of the form (Equation16).

Necessity of Item 1. From the assumption rank(uf)=2 (i.e. the assumption that the system has no redundant inputs), it immediately follows that dim(D1)=4, which in case of k1=1 already shows the necessity of item 1. For k12, the involutive distributions Δi can always be written in the form (Equation16) (see Remark 7.1), based on which dim(Di)=2(i+1), i=2,,k1 can be shown by contradiction. Assume that dim(Di)<2(i+1) for some i where 2ik1 and let 2lk1 be the smallest integer such that dim(Dl)<2(l+1). Thus adf¯lu¯1 and adf¯lu¯2 are collinear modulo Dl1. Both of these vector fields being contained in Dl1, i.e. adf¯lu¯1Dl1 and adf¯lu¯2Dl1, contradicts with Dk1 being non-involutive, it would lead to Dk1=Dl1, i.e. the sequence would stop growing from Dl1 on.

If adf¯lu¯2Dl1, we have (see (Equation16)) (18) Δl+1=span{u¯21,u¯11}+Dl1+span{adf¯l+1u¯2Dl},(18) where adf¯l+1u¯2Dl follows from the assumption adf¯lu¯2Dl1 since by construction Dl=Dl1+[f¯,Dl1]. Since Dl11Dl, we either have Δl+1=span{u¯21,u¯11}+Dl or Δl+1=span{u¯21,u¯11}+Dl1, i.e. the vector field adf¯l+1u¯2 in (Equation18) either completes Dl1 to Dl, or it is contained in Dl1. The case Δl+1=span{u¯21,u¯11}+Dl would lead to Δk1+1=span{u¯21,u¯11}+Dk1, which would imply that Dk1 is involutive, contradicting with Dk1 being non-involutive. Similarly, the case Δl+1=span{u¯21,u¯11}+Dl1 would lead to Δk1+2=span{u¯21,u¯11}+Dk1, which would again imply that Dk1 is involutive, again contradicting with Dk1 being non-involutive.

If adf¯lu¯2Dl1, we necessarily have Dl=Dl1+span{adf¯lu¯2} (since by assumption dim(Dl)<2(l+1) and dim(Dl1)=2l), which leads to Dk1=Dk11+span{adf¯k1u¯2}. From the involutivity of Δk1T(X×U)=Dk12+span{adf¯k11u¯2,adf¯k1u¯2}Dk1 it follows that [adf¯k11u¯2,adf¯k1u¯2]Dk1, which together with the involutivity of Dk11 and the fact that adf¯k11u¯2Dk11 implies that adf¯k11u¯2C(Dk1) (recall that Dk1=Dk11+span{adf¯k1u¯2}). From the Jacobi identity [adf¯k11u¯1,[f¯,adf¯k11u¯2]]+[adf¯k11u¯2C(Dk1),[adf¯k11u¯1,f¯]Dk1]Dk1+[f¯,[adf¯k11u¯2,adf¯k11u¯1]Dk11]Dk1=0it follows that [adf¯k11u¯1,adf¯k1u¯2]Dk1, which because of the involutivity of Dk11 implies that adf¯k11u¯1C(Dk1). Since also Dk12C(Dk1) (which can be shown based on the Jacobi identity), it follows that Dk11C(Dk1) which would imply that Dk1=Dk11+span{adf¯k1u¯2} is involutive, contradicting with Dk1 being non-involutive and completing the proof of the necessity of item 1.

In the following, we show the necessity of the remaining items. We distinguish between two cases, namely the case that the involutive distributions Δi in (Equation15) can be written in the form (Equation16) and the case that these distributions can only be written in the form (Equation17). As already mentioned, the form (Equation17) is only relevant if in the case k1=1 we have adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}, in all other cases, the distributions Δi are indeed of the form (Equation16). We treat this special case in Appendix 1 so that for the rest of the proof we can assume that the involutive distributions Δi in (Equation15) can be written in the form (Equation16).

Necessity of Item 2. We either have Dk11C(Dk1) or Dk11C(Dk1). We have to show the necessity of item 2a under the assumption that Dk11C(Dk1) and the necessity of item 2b under the assumption that Dk11C(Dk1).

The case Dk11C(Dk1)

We have to show the necessity of item 2a. We have Dk1Δk1+2T(X×U) (see (Equation16)) and since Δk1+2T(X×U) is involutive, it follows that D¯k1Δk1+2T(X×U). In turn, we either have dim(D¯k1)=dim(Dk1)+1, which corresponds to item 2a.A, or dim(D¯k1)=dim(Dk1)+2, which corresponds to 2a.B. We have to distinguish between these two possible subcases. In either subcase, the following result, proven in Appendix 2, will be useful.

Lemma 7.2

In the case d = 2 with Dk11C(Dk1) we have Dk1(1)=Dk1+span{adf¯k1+1u¯2}.

Let us first consider the subcase dim(D¯k1)=dim(Dk1)+1, which corresponds to item 2a.A. We have to show that dim([f¯,Dk1]+D¯k1)=dim(D¯k1)+1. We necessarily have D¯k1T(X×U), since D¯k1=T(X×U) would imply flatness with a difference of d = 1,Footnote3 which contradicts with the assumption that d = 2. Since D¯k1=Dk1(1)=Dk1+span{adf¯k1+1u¯2} (see Lemma 7.2), the condition holds if and only if adf¯k1+1u¯1D¯k1, which can be shown by contradiction. Assume that adf¯k1+1u¯1D¯k1 and thus [f¯,Dk1]+D¯k1=D¯k1. Based on the Jacobi identity it can then be shown that this would imply [f¯,D¯k1]D¯k1. However, since D¯k1T(X×U), this would in turn imply that the system contains an autonomous subsystem, which is in contradiction with the assumption that the system is flat.Footnote4 Thus adf¯k1+1u¯1D¯k1 and dim([f¯,Dk1]+D¯k1)=dim(D¯k1)+1 indeed holds. By definition we have Ek1+1=D¯k1 in this case, which evaluates to Ek1+1=Dk1+span{adf¯k1+1u¯2}.

Next, let us consider the subcase dim(D¯k1)=dim(Dk1)+2, which corresponds to item 2a.B. According to Lemma 7.2, we have Dk1(1)=Dk1+span{adf¯k1+1u¯2}, which is non-involutive since by assumption dim(D¯k1)=dim(Dk1)+2. Note that Dk1(1) can be written in the form Dk1(1)=Dk11+span{adf¯k1u¯2,adf¯k1+1u¯2}=Δk1+1T(X×U)+span{adf¯k1u¯1}.From the involutivity of Δk1+1T(X×U) and the fact that [adf¯k1u¯2,adf¯k1u¯1]Dk1(1) it follows that adf¯k1u¯2C(Dk1(1)). Based on the Jacobi identity, it can be shown that Dk11C(Dk1(1)). The non-involutivity of Dk1(1) implies that dim(C(Dk1(1)))dim(Dk1(1))2. We thus have C(Dk1(1))=Dk11+span{adf¯k1u¯2}, from which the necessity of the condition [f¯,C(Dk1(1))]Dk1(1) follows immediately. By definition we have Ek1+1=Dk1(1) in this case, which is of course non-involutive and evaluates to Ek1+1=Dk1+span{adf¯k1+1u¯2}.

The case Dk11C(Dk1)

We have to show the necessity of item 2b, i.e. we have to show that there exists a vector field vcDk11, vcDk12 such that the distributions Ek11=Dk12+span{vc} and Ek1=Dk11+span{[vc,f¯]} meet Ek11C(Ek1). Let us show that the vector field vc=adf¯k11u¯2 and the distributions Ek11 and Ek1 derived from it meet these criteria. The vector field vc=adf¯k11u¯2 meets adf¯k11u¯2Dk11, adf¯k11u¯2Dk12 (the latter must hold since adf¯k11u¯2Dk12 would imply dim(Dk11)<2k1), and yields Ek11=Dk12+span{adf¯k11u¯2} and Ek1=Dk11+span{adf¯k1u¯2}. Due to the involutivity of Dk11 and the involutivity of Δk1T(X×U)=Dk12+span{adf¯k11u¯2,adf¯k1u¯2}Ek1, it follows that [adf¯k11u¯2,Ek1]Ek1 and [Dk12,Ek1]Ek1, and hence Ek11C(Ek1). Therefore, vc=adf¯k11u¯2, Ek11=Dk12+span{adf¯k11u¯2} and Ek1=Dk11+span{adf¯k1u¯2}.

Necessity of Item 3. The distribution Ek1=Dk11+span{adf¯k1u¯2} of item 2b can either be involutive or non-involutive. We have to show the necessity of item 3a under the assumption that Ek1=Dk11+span{adf¯k1u¯2} of item 2b is non-involutive, and the necessity of item 3b under the assumption that Ek1 of item 2b is involutive.

In item 2a.A, the distribution Ek1+1 is by construction involutive, so we have to show the necessity of item 3b in this case. (In item 2a.B, the item 3 is not relevant as it is skipped in the corresponding conditions.)

Ek1 of item 2b being non-involutive

We have to show the necessity of item 3a. From Ek1Δk1+1T(X×U) (see (Equation16)), it immediately follows that E¯k1=Δk1+1T(X×U)=Dk11+span{adf¯k1u¯2,adf¯k1+1u¯2} and in turn the necessity of the condition dim(E¯k1)=dim(Ek1)+1, which corresponds to 3a.I follows.Footnote5

Next, let us show the necessity of item 3a.II, i.e. dim([f¯,Ek1]+E¯k1)=dim(E¯k1)+1. Recall that we have Ek1=Dk11+span{adf¯k1u¯2} and E¯k1=Dk11+span{adf¯k1u¯2,adf¯k1+1u¯2}. Since adf¯k1+1u¯2E¯k1, the condition holds if and only if adf¯k1u¯1E¯k1, which can be shown by contradiction. Assume that adf¯k1u¯1E¯k1 and thus [f¯,Ek1]+E¯k1=E¯k1. Based on the Jacobi identity, it can then be shown that this would imply [f¯,E¯k1]E¯k1. However, since E¯k1T(X×U), this would in turn imply that the system contains an autonomous subsystem, which is in contradiction with the assumption that the system is flat. By definition we have Fk1+1=E¯k1=Ek1+span{adf¯k1+1u¯2} in this case.

Ek1 of item 2b or Ek1+1 of item 2a.A being involutive

In this case, the distributions Ei=Ei1+[f¯,Ei1] are defined and we have to show the necessity of item 3b.

Let us first show the necessity of 3b.I, i.e. there necessarily exists an integer k2 such that Ek2 is non-involutive. We show the existence of k2 by contradiction. Assume that all the distributions Ei are involutive. If there does not exist an integer s such that Es=T(X×U), then there exists an integer l such that [f¯,El]El, implying that the system contains an autonomous subsystem and contradicting with the assumption that the system is flat. If all the distributions Ei are involutive and there exists an integer s such that Es=T(X×U), it can be shown that the system would meet the conditions for flatness with d = 1, which contradicts with the assumption that d = 2.

To show the condition on the dimensions of the distributions Ei, note that actually in any case, i.e. independently of whether 2a.A or 2b applies, we have Ek1+1=Dk1+span{adf¯k1+1u¯2}. Indeed, in 2a.A by definition we have Ek1+1=D¯k1, which turned out to be Ek1+1=Dk1+span{adf¯k1+1u¯2}. In 2b, we found that Ek1=Dk11+span{adf¯k1u¯2} (assumed to be involutive here) and thus again Ek1+1=Dk1+span{adf¯k1+1u¯2}. For the distributions Ei, k1+1ik2 we thus have (19) Ek1+1=span{u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯k1u¯1,adf¯k1u¯2,adf¯k1+1u¯2}Ek21=span{u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯k22u¯1,adf¯k22u¯2,adf¯k21u¯2}Ek2=span{u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯k21u¯1,adf¯k21u¯2,adf¯k2u¯2}(19) and with the distributions (Equation16), they are related via (20) Δk1+2=span{u¯21,u¯11}+Ek1+1+span{adf¯k1+2u¯2}Δk2=span{u¯21,u¯11}+Ek21+span{adf¯k2u¯2}Δk2+1=span{u¯21,u¯11}+Ek2+span{adf¯k2+1u¯2}.(20) The necessity of the condition dim(Ei)=2i+1, i=k1+1,,k2 can now be shown by contradiction. Assume that dim(Ei)2i for some i where k1+1ik2 and let k1+1lk2 be the smallest integer such that dim(El)2l. We have El1=span{u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯l2u¯1,adf¯l2u¯2,adf¯l1u¯2}El=span{u¯1,u¯2,[f¯,u¯1],[f¯,u¯2],,adf¯l1u¯1,adf¯l1u¯2,adf¯lu¯2}and by assumption dim(El1)=2l1 and dim(El)2l. Thus adf¯l1u¯1 and adf¯lu¯2 are collinear modulo El1. Both of these vector fields being contained in El1, contradicts with Ek2 being non-involutive, it would lead to Ek2=El1, i.e. the sequence would stop growing from El1 on. If adf¯lu¯2El1, then due to Δl=span{u¯21,u¯11}+El1+span{adf¯lu¯2}(see (Equation20)), we have Δl=span{u¯21,u¯11}+El1 and it follows that also Δi=span{u¯21,u¯11}+Ei1 for i>l. The involutivity of Δk2+1 would then imply that Ek2 is involutive, which again contradicts with Ek2 being non-involutive.

If adf¯u¯2El1, we necessarily have adf¯l1u¯1El1+span{adf¯lu¯2} and thus El=El1+span{adf¯lu¯2} and in turn Δl=span{u¯21,u¯11}+El and it follows that also Δi=span{u¯21,u¯11}+Ei for i>l. The involutivity of Δk2 would then imply that Ek2 is involutive, which again contradicts with Ek2 being non-involutive. Thus adf¯l1u¯1 and adf¯lu¯2 cannot be collinear modulo El1 for any k1+1lk2 which shows that dim(Ei)=2i+1 for i=k1+1,,k2.Footnote6

Necessity of Item 4. For the distributions Ek21 and Ek2 of item 3b, we either have Ek21C(Ek2) or Ek21C(Ek2). We have to show the necessity of item 4a under the assumption that Ek21C(Ek2), and the necessity of item 4b under the assumption that Ek21C(Ek2). Furthermore, under the assumption that the conditions of 2a.B are met, we have to show the necessity of item 4a.II.

The case Ek21C(Ek2)

We have to show the necessity of item 4a. The distribution Ek2 is by assumption non-involutive. Since Ek2Δk2+1T(X×U) (see (Equation20)) and Δk2+1T(X×U) is involutive, it follows that E¯k2=Δk2+1T(X×U)=Ek2+span{adf¯k2+1u¯2}, and thus we have dim(E¯k2)=dim(Ek2)+1, which shows the necessity 4a.I.

For E¯k2=T(X×U), there are no additional conditions whose necessity has to be shown. For E¯k2T(X×U), the necessity of dim([f¯,Ek2]+E¯k2)=dim(E¯k2)+1 has to be shown. Since adf¯k2+1u¯2E¯k2, the condition holds if and only if adf¯k2u¯1E¯k2, which can be shown by contradiction. Assume that adf¯k2u¯1E¯k2 and thus [f¯,Ek2]+E¯k2=E¯k2. Based on the Jacobi identity, it can then be shown that this would imply [f¯,E¯k2]E¯k2, which would in turn imply that the system contains an autonomous subsystem and contradict with the system being flat. By definition we have Fk2+1=E¯k2=Ek2+span{adf¯k2+1u¯2} in this case.

The necessity of item 4a.II under the assumption that the conditions of 2a.B are met can be shown analogously. By definition we have Ek2=Dk1(1) (where k2=k1+1), which according to Lemma 7.2 evaluates to Ek2=Dk1+span{adf¯k1+1u¯2} and is by assumption non-involutive. We have Ek2Δk1+2T(X×U) (see (Equation16)), from which E¯k2=Δk1+2T(X×U)=Ek2+span{adf¯k1+2u¯2} follows. For E¯k2=T(X×U), there are again no additional conditions whose necessity has to be shown. For E¯k2T(X×U), the necessity of dim([f¯,Ek2]+E¯k2)=dim(E¯k2)+1 can be shown analogously as above. By definition we have Fk2+1=E¯k2=Ek2+span{adf¯k2+1u¯2} in this case.

The case Ek21C(Ek2)

We have to show the necessity of item 4b, i.e. we have to show that the distribution Fk2, defined as Fk2=Ek21+C(Ek2), is involutive. Regarding the Cauchy characteristic distribution of Ek2 we have the following result, a proof of which is provided in Appendix 2.

Lemma 7.3

Assume that Ek21C(Ek2). If k2k1+2, the Cauchy characteristic distribution of Ek2 is given by C(Ek2)=Ek22+span{adf¯k21u¯2,adf¯k2u¯2λadf¯k22u¯1},and if k2=k1+1, it is given by C(Ek2)=Δk1Dk11+span{adf¯k21u¯2,adf¯k2u¯2λadf¯k22u¯1},in both cases with some function λ(x,u¯).

Recall that we have Ek21=span{u¯1,u¯2,,adf¯k22u¯1,adf¯k22u¯2,adf¯k21u¯2} and that Dk11Ek1. Due to Lemma 7.3, for the distribution Fk2=Ek21+C(Ek2) we thus in any case have Fk2=Ek21+span{adf¯k2u¯2} and hence Fk2=Δk2T(X×U), implying that Fk2 is indeed involutive.

Necessity of Item 5. In conclusion, in 3a, i.e. for Ek1 being non-involutive (in which case we define k2=k1), we have Fk1+1=E¯k1=Ek1+span{adf¯k1+1u¯2}. In 4a, we have Fk2+1=E¯k2=Ek2+span{adf¯k2+1u¯2}, and in 4b, we have Fk2=Ek21+span{adf¯k2u¯2}. With Fi=Fi1+[f¯,Fi1], in any case we thus have Fk2+1=Ek2+span{adf¯k2+1u¯2}. To complete the necessity part of the proof we have to show that all the distributions Fi, ik2+1 are involutive and that there exists an integer s such that Fs=T(X×U), which is indeed the case since in any case it follows that the distributions Fi and Δi are related via Δi=span{u¯21,u¯11}+Fi and thus Fk2+1=Δk2+1T(X×U)Fk2+2=Δk2+2T(X×U)Fs=ΔsT(X×U)=T(X×U).

Sufficiency

Consider a two-input system of the form (Equation1) and assume that it meets the conditions of Theorem 5.1. To cover all the possible paths which are illustrated in Figure , we again have to distinguish between several cases. The distinction is done based on the difference of the integers k1 and k2, i.e. the indices of the first non-involutive distribution Dk1 and the second non-involutive distribution Ek2. The cases in which k2k1+2 and thus Ek1+1 is involutive are similar and can be proven together. The cases in which Ek1 is involutive but Ek1+1 is non-involutive are also similar and can be proven together. The remaining case in which Ek1 is non-involutive (in which case we define k2=k1) is proven separately. For each case, a coordinate transformation such that the system takes a certain structurally flat triangular form can be derived. For the cases k2=k1+1 and k2=k1 we derive such a transformation explicitly. The case k2k1+2 can be handled analogously, it is in fact slightly simpler than the other two cases and we do not treat this case in detail here. In each case, we will need the following result, proven in Appendix 2.

Lemma 7.4

Let Gk1 be an involutive distribution and Gk a non-involutive distribution on X×U such that Gk12Gk, Gk1C(Gk) and dim(G¯k)=dim(Gk)+1 and such that for some vector field f on X×U we have Gk=Gk1+[f,Gk1] and either G¯k=T(X×U) or dim([f,Gk]+G¯k)=dim(G¯k)+1. Then, there exists an involutive distribution Hk such that Gk11Hk1Gk and Hk+[f,Hk]=G¯k. In case of G¯kT(X×U), the distribution Hk is unique.

The case k2k1+2: In total, four different subcases are possible, namely

  1. 2a.A followed by 4a, corresponding to Dk11C(Dk1) and Ek21C(Ek2).

  2. 2a.A followed by 4b, corresponding to Dk11C(Dk1) and Ek21C(Ek2).

  3. 2b followed by 4a, corresponding to Dk11C(Dk1) and Ek21C(Ek2).

  4. 2b followed by 4b, corresponding to Dk11C(Dk1) and Ek21C(Ek2).

In any of these cases, we have the following sequence of involutive distributions: (21) D022Dk111Ek12Ek1+1=D¯k122Ek211Fk22Fk2+1=E¯k2Fs.(21) In the cases in which Dk11C(Dk1) and/or Ek21C(Ek2), the existence of Ek1 and/or Fk2 is guaranteed by Lemma 7.4. That the dimensions of the distributions in the inclusion Fk22Fk2+1 indeed differ by two follows from Fk21Ek21E¯k2=Fk2+1 in this case.

In case of Dk11C(Dk1), the distribution Ek1 occurs explicitly in the corresponding conditions of Theorem 5.1 and Ek1+1=D¯k1 follows from the assumption that Ek1+1 is involutive in this case. Indeed, we have Ek1+1=Ek1+[f,Ek1]=Dk1+span{[f,[vc,f]]}, from which D¯k1=Ek1+1 follows. Similarly, in case of Ek21C(Ek2), Fk2+1=E¯k2 can be shown as follows. By construction, we have Fk2=Ek21+C(Ek2) and Fk2+1=Fk2+[f,Fk2], which are by assumption involutive and Ek2 is non-involutive. It is immediate that C(Ek2) necessarily contains a vector field vcEk2, vcEk21, since otherwise we would have Fk2=Ek21, leading to Fk2+1=Ek2 and contradicting with Ek2 being non-involutive. On the other hand, C(Ek2) can only contain one vector field which is not already contained in Ek21, since otherwise Fk2=Ek2 which again contradicts with Ek2 being non-involutive. Thus Fk2=Ek21+span{vc} where vcEk2, vcEk21. Therefore, we have Fk2+1=Ek2+span{[f,vc]}. It is immediate that [f,vc]Ek2 (otherwise Fk2+1=Ek2). Thus we have E¯k2=Fk2+1. That Fk22Fk2+1 follows readily from these considerations in this case.

Based on the distributions (Equation21), a coordinate transformation such that in the new coordinates the system takes the structurally flat triangular form x˙s=fs(xs,xs1)x˙k2+2=fk2+2(xs,,xk2+1)x˙k2+1=fk2+1(xs,,xk21,xk212)x˙k21=fk21(xs,,xk21)x˙k21=fk21(xs,,xk22)x˙k1+2=fk1+2(xs,,xk1+1)x˙k1+1=fk1+1(xs,,xk11,xk112)x˙k11=fk11(xs,,xk11)x˙k11=fk11(xs,,xk12)x˙2=f2(xs,,x1)x˙1=f1(xs,,x1,u),can be derived, where dim(xi)=2, i{1,,k11,k1+1,,k21,k2+1}, dim(xk11)=dim(xk21)=1 (in case of k1=1 the variables xk11 are actually inputs instead of states). For dim(xs)=2 it follows that y=xs forms a flat output with a difference of d = 2. If dim(xs)=1, we necessarily have dim(xl)=2 but dim(fl+1)=1 for some l{k2+1,,s1}. In this case, a flat output with a difference of d = 2 is given by y=(xs,φ) where φ=φ(xs,,xl) is an arbitrary function such that rank(xl(fl+1,φ))=2. Furthermore, it follows that there cannot exist a flat output with d1. The conditions of Theorem 3.2 cannot be met due to the non-involutivity of Dk1 and it can easily be shown that the conditions of Theorem 5.2 cannot be met either.

The case k2=k1+1: In total, three different subcases are possible, namely

  1. 2a.B followed by 4a.II, corresponding to Dk11C(Dk1).

  2. 2b followed by 4a, corresponding to Dk11C(Dk1) and Ek1C(Ek1+1).

  3. 2b followed by 4b, corresponding to Dk11C(Dk1) and Ek1C(Ek1+1).

In all of these cases, we have the following sequence of involutive distributions: (22) D022Dk111Ek11Fk1+12Fk1+2=E¯k1+1Fk1+3Fs.(22) Let us show this in detail for the three possible cases (a)–(c).

(a) In 2a.B, only Ek1+1 is defined explicitly. We set Ek1=C(Dk1(1)) in this case. We have the following results on this Cauchy characteristic distribution, see Appendix 2 for a proof.

Lemma 7.5

Assume that the conditions of item 2a.B are met, i.e. Dk11C(Dk1), dim(D¯k1)=dim(Dk1)+2 and [f,C(Dk1(1))]Dk1(1). Then, Dk111C(Dk1(1))1Dk1 and C(Dk1(1))+[f,C(Dk1(1))]=Dk1(1).

By definition, we have Ek1+1=Dk1(1) and according to Lemma 7.5 we have Ek1+1=Ek1+[f,Ek1] with Ek1=C(Dk1(1)). Therefore, Lemma 7.4 applies, which guarantees the existence of a distribution Fk1+1 such that Ek11Fk1+11Ek1+1 and Fk1+1+[f,Fk1+1]=E¯k1+1 (Fk1+1 is unique if and only if D¯k1T(X×U)).

(b) In 2b, the distribution Ek1 occurs explicitly, the existence of an involutive distribution Fk1+1 such that Ek11Fk1+11Ek1+1 and Fk1+1+[f,Fk1+1]=E¯k1+1 is guaranteed by Lemma 7.4 (Fk1+1 is unique if and only if E¯k1+1T(X×U)).

(c) In 2b and 4b, the distributions Ek1 and Fk1+1 occur explicitly. That dim(Fk1+1)=dim(Ek1)+1 as well as Fk1+2=E¯k1+1 and dim(Fk1+2)=dim(Fk1+1)+2 can be shown the same way as above for the case k2k1+2.

Based on the distributions (Equation22), we will derive a coordinate transformation such that in the new coordinates the system takes the structurally flat triangular form (23) x˙s=fs(xs,xs1)x˙k1+3=fk1+3(xs,,xk1+2)x˙k1+2=fk1+2(xs,,xk1+11,xk112)x˙k1+11=fk1+11(xs,,xk11,xk112)x˙k11=fk11(xs,,xk11)x˙k11=fk11(xs,,xk12)x˙2=f2(xs,,x1)x˙1=f1(xs,,x1,u),(23) where dim(xi)=2, i{1,,k11,k1+2}, dim(xk11)=dim(xk1+11)=1 (in case of k1=1 the variables xk11 are actually inputs instead of states) from which again flatness with a difference of d = 2 can be deduced. For dim(xs)=2 it follows that y=xs forms a flat output with a difference of d = 2. If dim(xs)=1, we necessarily have dim(xl)=2 but dim(fl+1)=1 for some l{k1+2,,s1}. In this case, a flat output with a difference of d = 2 is given by y=(xs,φ) where φ=φ(xs,,xl) is an arbitrary function such that rank(xl(fl+1,φ))=2.

The case k2=k1: In this case, by assumption the conditions of item 2b and item 3a are met. We have the following sequence of involutive distributions (we assume k12 here, see Remark 7.4 below for the cases k1=1) (24) D022Dk121Ek111Fk12Fk1+1=E¯k1Fk1+2Fs.(24) That Ek11 is involutive follows from the fact that Ek11=C(Ek1) in this case. (Indeed, since Ek1 is non-involutive we necessarily have dim(C(Ek1))dim(Ek1)2. By construction we have Ek11C(Ek1) and Ek112Ek1, which thus implies Ek11=C(Ek1).) The existence of Fk1 such that Ek111Fk11Ek1 and Fk1+[f,Fk1]=E¯k1 is guaranteed by Lemma 7.4 (and Fk1 is unique since E¯k1=T(X×U) is not possible as it would imply Dk1=T(X×U)). Based on this sequence, we will derive a coordinate transformation such that in the new coordinates the system takes the structurally flat triangular form (25) x˙s=fs(xs,xs1)x˙k1+2=fk1+2(xs,,xk1+1)x˙k1+1=fk1+1(xs,,xk11,xk122)x˙k11=fk11(xs,,xk111,xk122)x˙k111=fk111(xs,,xk12)x˙k12=fk12(xs,,xk13)x˙2=f2(xs,,x1)x˙1=f1(xs,,x1,u),(25) where dim(xi)=2, i{1,,k12,k1+1}, dim(xk111)=dim(xk11)=1 (in case of k1=2 the variables xk12 are actually inputs instead of states) from which again flatness with a difference of d = 2 can be deduced. For dim(xs)=2 it follows that y=xs forms a flat output with a difference of d = 2. If dim(xs)=1, we necessarily have dim(xl)=2 but dim(fl+1)=1 for some l{k1+1,,s1}. In this case, a flat output with a difference of d = 2 is given by y=(xs,φ) where φ=φ(xs,,xl) is an arbitrary function such that rank(xl(fl+1,φ))=2.

The case k2=k1+1 in detail: Apply a change of coordinates x¯=Φx(x) such that all the distributions (Equation22) get straightened out simultaneously, i.e. such that Di=span{u¯,x¯1,,x¯i},i=0,,k11,where dim(x¯i)=2, and Ek1=span{u¯,x¯1,,x¯k11,x¯k11},where dim(x¯k11)=1, and Fi=span{u¯,x¯1,,x¯k11,x¯k11,x¯k1+11,x¯k1+2,,x¯i},i=k1+1,,s,where dim(x¯k1+11)=1 and dim(x¯k1+2)=2. Since by construction we have [f,Di]Di+1 for i=0,,k12 as well as [f,Fi]Fi+1 for i=k1+1,,s1, it follows that in the new coordinates the vector field f has the triangular form f=[fs(x¯s,x¯s1)fk1+3(x¯s,,x¯k1+2)fk1+2(x¯s,,x¯k11)fk1+11(x¯s,,x¯k11)fk11(x¯s,,x¯k11)fk11(x¯s,,x¯k12)f2(x¯s,,x¯1)f1(x¯s,,x¯1,u)].From Di+1=Di+[f,Di], i=0,,k12 and Fi+1=Fi+[f,Fi], i=k1+1,,s1, it follows that the rank conditions rank(uf1)=dim(f1)=2, rank(x¯ifi+1)=dim(fi+1)=2 for i=1,,k12 and rank(x¯ifi+1)=dim(fi+1) for i=k1+2,,s1, hold.

Since Dk112Dk1, it follows that rank(x¯k11(fk1+2,fk1+11,fk11))=2. The non-involutivity of Dk1 implies that at least one of the components of fk1+2 must explicitly depend on x¯k11 (otherwise Dk1 would be involutive as it would be straightened out in the x¯ coordinates). Without loss of generality we can thus assume that fk1+22 explicitly depends on x¯k112, (if not, permute x¯k111 and x¯k112 and/or x¯k1+21 and x¯k1+22). We can thus introduce x~k112=fk1+22 with all the other coordinates left unchanged (in case of k1=1 this is actually an input transformation instead of a state transformation), resulting in f~k1+2=[x~k112f~k1+21(x¯s,,x~k11)],(where x~k11=(x~k112,x¯k111)). In any case, we have Ek11Dk1.Footnote7 Thus, a linear combination of [x¯k111,f] and [x~k112,f] is contained in Ek1, implying that rank(x~k11(f~k1+2,f~k1+11))=1 and thus, f~k1+21 and f~k1+11 are actually independent of x¯k111, so we have f~k1+2=[x~k112f~k1+21(x¯s,,x¯k11,x~k112)]and f~k1+11=f~k1+11(x¯s,,x¯k11,x~k112).Furthermore, we have Ek12Ek1+1 and thus rank((x¯k11,x~k112)(f~k1+2,f~k1+11))=2 and since Fk1+11Ek1+1, a linear combination of [x¯k11,f] and [x~k112,f] is contained in Fk1+1, implying that rank((x¯k11,x~k112)f~k1+2)=1 and thus, f~k1+21 is actually independent of x¯k11, so we have f~k1+2=[x~k112f~k1+21(x¯s,,x¯k1+11,x~k112)]and f~k1+11 must explicitly depend on x¯k11. Because of Fk1+12Fk1+2 we have rank((x~k112,x¯k1+11)f~k1+2)=2, from which it follows that f~k1+21 explicitly depends on x¯k1+11. (The non-involutivity of Ek1+1 furthermore implies that x~k1122f~k1+210 and/or x¯k1+11x~k112f~k1+210.) So in the just constructed coordinates, the system equations indeed take the triangular structure (Equation23).

Remark 7.2

In case (a), i.e. if Dk11C(Dk1), by successively introducing new coordinates in (Equation23) from top to bottom, the system would take the triangular from proposed in Gstöttner et al. (Citation2021a). The cases (b) and (c) are not covered by the results in Gstöttner et al. (Citation2021a).

The case k2=k1 in detail: Apply a change of coordinates x¯=Φx(x) such that all the distributions (Equation24) get straightened out simultaneously, i.e. such that Di=span{u¯,x¯1,,x¯i},i=0,,k12,where dim(x¯i)=2, and Ek11=span{u¯,x¯1,,x¯k12,x¯k111},where dim(x¯k111)=1, and Fi=span{u¯,x¯1,,x¯k12,x¯k111,x¯k11,x¯k1+1,,x¯i},i=k1,,s,where dim(x¯k11)=1 and dim(x¯k1+1)=2. Since by construction we have [f,Di]Di+1 for i=0,,k13 as well as [f,Fi]Fi+1 for i=k1,,s1, it follows that in the new coordinates the vector field f has the triangular form (26) f=[fs(x¯s,x¯s1)fk1+2(x¯s,,x¯k1+1)fk1+1(x¯s,,x¯k12)fk11(x¯s,,x¯k12)fk111(x¯s,,x¯k12)fk12(x¯s,,x¯k13)f2(x¯s,,x¯1)f1(x¯s,,x¯1,u)].(26) From Di+1=Di+[f,Di], i=0,,k13 and Fi+1=Fi+[f,Fi], i=k1,,s1, it follows that the rank conditions rank(uf1)=dim(f1)=2, rank(x¯ifi+1)=dim(fi+1)=2 for i=1,,k13 and rank(x¯ifi+1)=dim(fi+1) for i=k1+1,,s1, hold.

It follows that at least one component of fk1+1 explicitly depends on x¯k12. Indeed, we have Dk122Dk11. If fk1+1 would be independent of x¯k12, we would obtain Dk11=Dk12+span{x¯k111,x¯k11}=Fk1 and in turn Dk1=Fk1+1, which contradicts with Dk1 being non-involutive. Without loss of generality, we can thus assume that fk1+12 explicitly depends on x¯k122 (if not, permute x¯k121 and x¯k122 and/or x¯k1+11 and x¯k1+12). We can thus introduce x~k122=fk1+12 with all the other coordinates left unchanged (in case of k1=2 this is actually an input transformation instead of a state transformation), resulting in f~k1+1=[x~k122f~k1+11(x¯s,,x~k12)](where x~k12=(x~k122,x¯k121)). Since Fk11Ek1, it follows that [f,Ek11] yields exactly one new direction which is not already contained in Fk1 and thus, rank((x¯k111,x~k122,x¯k121)f~k1+1)=1 and therefore, f~k1+11 is actually independent of x¯k121 and x¯k111, i.e. we actually have f~k1+1=[x~k122f~k1+11(x¯s,,x¯k11,x~k122)].Since Fk12Fk1+1, it follows that fk1+11 explicitly depends on x¯k11. Similarly, we have Ek111Dk11, from which it follows that [f,Dk12] yields exactly one new direction which is not already contained in Ek11 and thus, rank((x~k122,x¯k121)(f~k1+1,f~k11))=1 and therefore, f~k11 is actually independent of x¯k121, i.e. we actually have f~k11=f~k11(x¯s,,x¯k111,x~k122).Since Ek112Ek1, it follows that f~k11 explicitly depends on x¯k111. So in the just constructed coordinates, the system equations indeed take the triangular structure (Equation25). (Note that the involutivity of Dk11 implies that the functions f~k1+11 and f~k11 depend on x~k122 in an affine manner and that furthermore x¯k111x~k122f~k11=0. The non-involutivity of Ek1 implies that x¯k11x~k122f~k1+110.)

Remark 7.3

By successively introducing new coordinates in (Equation25) from top to bottom, the system would take the triangular from proposed in Gstöttner et al. (Citation2020b).

Remark 7.4

In case of k1=1, the variables x¯k11 would correspond to inputs of the system and the variables x¯k12 would not exist. Consider the system obtained by one-fold prolonging both of its inputs, i.e.  x˙=f(x,u)u˙1=u11u˙2=u12with the state (x,u1,u2) and the input (u11,u12). By assumption, the original system meets the conditions of Theorem 5.1 with k1=k2=1. It can easily be shown that this implies that the prolonged system also meets the conditions of Theorem 5.1, but with k1=k2=2. (The distributions involved in the conditions of Theorem 5.1 when applying it to the original system and when applying it to the prolonged system differ only by span{u11,u12}.) The prolonged system can thus be transformed into the corresponding triangular form (Equation25) as explained above, i.e. the prolonged system can be proven to be flat with a difference of d = 2. It can be shown that the prolonged system is flat with a certain difference d if and only if the original system is flat with the same difference d. In fact, a flat output of the original system with a certain difference d is also a flat output of the prolonged system with the same difference d and vice versa. The prolonged system being flat with a difference of d = 2 thus implies that the original system is flat with a difference of d = 2.

8. Brief sketch of the proof of Theorem 5.2

Necessity

Consider a two-input system of the form (Equation1) and assume that it is flat with a difference of d = 1. According to Theorem 3.3, there exists an input transformation u¯=Φu(x,u) with inverse u=Φˆ(x,u¯) such that the system obtained by onefold prolonging the new input u¯1, i.e. the system (27) x˙=f(x,Φˆu(x,u¯))=f¯(x,u¯)u¯˙1=u¯11(27) with the state (x,u¯1) and the input (u¯11,u¯2), is static feedback linearisable. The necessity of the conditions of Theorem 5.2 can then be shown on basis of the involutive distributions Δi=Δi1+[fp,Δi1], where Δ0=span{u¯11,u¯2} and fp=f¯i(x,u¯)xi+u¯11u¯1, which are involved in the test for static feedback linearisability of the prolonged system (see Theorem 3.2).

The necessity part of the proof is in fact very similar to the proof of the necessity of the items 3b to 5 of Theorem 5.1. As we have already noted above, these items in fact coincide with the items 1 to 3 of Theorem 5.2 when Ei and k2 are replaced by Di and k1.

Sufficiency

Consider a two-input system of the form (Equation1) and assume that it meets the conditions of Theorem 5.2. In any of the two cases, i.e. independent of Dk11C(Dk1) (which corresponds to 2a) or Dk11C(Dk1) (which corresponds to 2b), it follows that we have the following sequence of involutive distributions: (28) D02D122Dk111Ek12Ek1+1=D¯k1Es=T(X×U),(28) based on which a change of coordinates such that in the new coordinates the system takes the structurally flat triangular form (29) x˙s=fs(xs,xs1)x˙k1+2=fk1+2(xs,,xk1+1)x˙k1+1=fk1+1(xs,,xk11,xk112)x˙k11=fk11(xs,,xk11)x˙k11=fk11(xs,,xk12)x˙2=f2(xs,,x1)x˙1=f1(xs,,x1,u),(29) can be derived (in case of k1=1, the variables xk11 are actually inputs instead of states), from which it follows that the system is indeed flat with a difference of d = 1.

Remark 8.1

In case of 2a, i.e. if Dk11C(Dk1), by successively introducing new coordinates in (Equation29) from top to bottom, the system would take the triangular from proposed in Gstöttner et al. (Citation2021a).

9. Conclusion

We have derived necessary and sufficient conditions for the linearisability of two-input systems by an endogenous dynamic feedback with a dimension of at most two. Verifying these conditions requires differentiation and algebraic operations only. Further research will be devoted to two-input systems which are linearisable by a three-dimensional endogenous dynamic feedback. Furthermore, we aim to extend results of this paper to systems with more than two inputs in future work. The class of two-input systems linearisable by an at most two-dimensional endogenous dynamic feedback overlaps with several other classes of flat systems which have been considered in the literature, e.g. the class of flat two-input driftless systems (see Martin and Rouchon (Citation1994)) or the class of flat control affine systems with four states and two inputs considered in Pomet (Citation1997). Elaborating in detail the relation of the conditions for flatness for these classes of systems with the conditions we have proposed in this paper also gives rise to future work.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Correction Statement

This article has been republished with minor changes. These changes do not impact the academic content of the article.

Additional information

Funding

The first author has been supported by the Austrian Science Fund (FWF) under grant number P 32151.

Notes

1 There exist at most two distinct such pairs of distributions Ek11 and Ek1, the construction is explained below. If indeed two pairs exist, a branching point occurs and we have to continue with both of them.

2 We have k1=1 and D0C(D1). The conditions of Theorem 5.2 are not met since E1 of item 2b is non-involutive. The conditions of Theorem 5.1 are not met since dim(E¯1)=dim(E1)+2, which violates 3a.I.

3 Indeed, for Dk11C(Dk1) with dim(D¯k1)=dim(Dk1)+1 and D¯k1=T(X×U), the conditions of Theorem 5.2 would be met.

4 The autonomous subsystems occurs explicitly in coordinates in which the distribution D¯k1 is straightened out (see also Theorem 3.49 in Nijmeijer and van der Schaft (Citation1990)).

5 Note that we necessarily have E¯k1T(X×U). Indeed, we have dim(Ek1)=dim(Dk1)1, and we have just shown that necessarily dim(E¯k1)=dim(E¯k1)+1. Thus, dim(E¯k1)=dim(Dk1) and E¯k1=T(X×U) would thus imply Dk1=T(X×U), which is in contradiction with Dk1 being non-involutive.

6 In case of 2a.A, there does not explicitly occur a distribution Ek1, so for l=k1+1, we cannot argue as above. However, in this case we have Ek1+1=D¯k1 and therefore, dim(Ek1+1)=2(k1+1)+1 follows immediately from dim(Dk1)=2(k1+1) and dim(D¯k1)=dim(Dk1)+1.

7 In 2b, this follows immediately from the definition of Ek1, for the case 2a.B, it follows from Lemma 7.5.

8 Note that we have [v2,[v1,f]]=[v1,[v2,f]] mod Dk1.

9 It is immediate that adf¯k1+1Dk1, as it would either lead to Δk1+1T(X×U)=Dk1 or Δk1+2T(X×U)=Dk1, which contradicts with Dk1 being non-involutive.

10 Note that there is no need to explicitly construct the vector fields vj=vji2(z2)z2i2, j = 1, 2 for deriving the distribution Hk corresponding to a certain choice ψ. Indeed, any vector fields w1,w2 which complete Gk1 to Gk can be written as a linear combination wj=βj1v1+βj2v2 mod Gk1 with some functions βji=βji(z2,z1) and β11β22β12β210. A straight forward calculation shows that (dψw2)w1(dψw1)w2=(β11β22β12β21)((dψv2)v1(dψv1)v2) mod Gk1 and thus H~k=Gk1+span{(dψw2)w1(dψw1)w2}=Gk11+span{(dψv2)v1(dψv1)v2}=Hk, i.e. the particular choice for the vector fields w1,w2 has no effect as long as they complete Gk1 to Gk.

11 It follows that the direction of the vector field v=v2αv1 is modulo Gk1 uniquely determined by the conditions vGk, vGk1, [v,f]G¯kT(X×U) and thus, the distribution Hk is unique in this case. Indeed, assume that there would exist another such vector field wGk, wGk1, which is modulo Gk1 not collinear with v and still satisfies [w,f]G¯k. Then, we would have Gk=Gk1+span{v,w} and in turn [f,Gk]G¯k, which is in contradiction with dim([f,Gk]+G¯k)=dim(G¯k)+1.

References

  • Fliess, M., Lévine, J., Martin, P., & Rouchon, P. (1992). Sur les systèmes non linéaires différentiellement plats. Comptes rendus de l'Académie des sciences. Série I, Mathématique, 315, 619–624.
  • Fliess, M., Lévine, J., Martin, P., & Rouchon, P. (1995). Flatness and defect of non-linear systems: Introductory theory and examples. International Journal of Control, 61(6), 1327–1361. https://doi.org/10.1080/00207179508921959
  • Fliess, M., Lévine, J., Martin, P., & Rouchon, P. (1999). A Lie–Bäcklund approach to equivalence and flatness of nonlinear systems. IEEE Transactions on Automatic Control, 44(5), 922–937. https://doi.org/10.1109/9.763209
  • Fossas, E., & Franch, J. (2000). Linearization by prolongations and quotient of modules, a case study: The vertical take off and landing aircraft. In Proceedings wses international conference on applied and theoretical mathematics (pp. 1961–1965).
  • Gstöttner, C., Kolar, B., & Schöberl, M. (2020a). On the linearization of flat two-input systems by prolongations and applications to control design. IFAC-PapersOnLine, 53(2), 5479–5486. 21st IFAC World Congress. https://doi.org/10.1016/j.ifacol.2020.12.1553
  • Gstöttner, C., Kolar, B., & Schöberl, M. (2020b). A structurally flat triangular form based on the extended chained form. International Journal of Control. https://doi.org/10.1080/00207179.2020.1841302
  • Gstöttner, C., Kolar, B., & Schöberl, M. (2021a). On a flat triangular form based on the extended chained form. IFAC-PapersOnLine, 54(9), 245–252. 24th International symposium on mathematical theory of networks and systems (MTNS). https://doi.org/10.1016/j.ifacol.2021.06.082
  • Gstöttner, C., Kolar, B., & Schöberl, M. (2021b). A finite test for the linearizability of two-input systems by a two-dimensional endogenous dynamic feedback. arXiv:2011.06929 [math.DS]. ECC 2021.
  • Jakubczyk, B., & Respondek, W. (1980). On linearization of control systems. Bulletin L'Académie Polonaise des Science, Série des Sciences Mathématiques, 28, 517–522.
  • Kai, T. (2006). Extended chained forms and their application to nonholonomic kinematic systems with affine constraints: control of a coin on a rotating table. In Proceedings of the 45th IEEE conference on decision and control (pp. 6104–6109). https://doi.org/10.1109/CDC.2006.377186
  • Kolar, B., Schöberl, M., & Schlacher, K. (2016). Properties of flat systems with regard to the parameterization of the system variables by the flat output. IFAC-PapersOnLine, 49(18), 814–819. 10th IFAC symposium on nonlinear control systems NOLCOS 2016. https://doi.org/10.1016/j.ifacol.2016.10.266
  • Lévine, J. (2009). Analysis and control of nonlinear systems: a flatness-based approach. Springer.
  • Li, S., Nicolau, F., & Respondek, W. (2016). Multi-input control-affine systems static feedback equivalent to a triangular form and their flatness. International Journal of Control, 89(1), 1–24. https://doi.org/10.1080/00207179.2015.1056232
  • Li, S., Xu, C., Su, H., & Chu, J. (2013). Characterization and flatness of the extended chained system. In Proceedings of the 32nd Chinese control conference (pp. 1047–1051). IEEE.
  • Martin, P., Devasia, S., & Paden, B. (1996). A different look at output tracking: control of a vtol aircraft. Automatica, 32(1), 101–107. https://doi.org/10.1016/0005-1098(95)00099-2
  • Martin, P., & Rouchon, P. (1994). Feedback linearization and driftless systems. Mathematics of Control, Signals and Systems, 7(3), 235–254. https://doi.org/10.1007/BF01212271
  • Nicolau, F., & Respondek, W. (2016a, December). Flatness of two-input control-affine systems linearizable via a two-fold prolongation. In 2016 IEEE 55th conference on decision and control (CDC) (pp. 3862–3867). https://doi.org/10.1109/CDC.2016.7798852
  • Nicolau, F., & Respondek, W. (2016b). Two-input control-affine systems linearizable via one-fold prolongation and their flatness. European Journal of Control, 28, 20–37. https://doi.org/10.1016/j.ejcon.2015.11.001
  • Nicolau, F., & Respondek, W. (2017). Flatness of multi-input control-affine systems linearizable via one-fold prolongation. SIAM Journal on Control and Optimization, 55(5), 3171–3203. https://doi.org/10.1137/140999463
  • Nicolau, F., & Respondek, W. (2019). Normal forms for multi-input flat systems of minimal differential weight. International Journal of Robust and Nonlinear Control, 29(10), 3139–3162. https://doi.org/10.1002/rnc.v29.10
  • Nicolau, F., & Respondek, W. (2020). Normal forms for flat two-input control systems linearizable via a two-fold prolongation. IFAC-PapersOnLine, 53(2), 5441–5446. 21st IFAC World Congress. https://doi.org/10.1016/j.ifacol.2020.12.1545
  • Nijmeijer, H., & van der Schaft, A. (1990). Nonlinear dynamical control systems. Springer.
  • Pomet, J. (1997). On dynamic feedback linearization of four-dimensional affine control systems with two inputs. ESAIM: Control, Optimisation and Calculus of Variations, 2(1997), 151–230. https://doi.org/10.1051/cocv:1997107
  • Schlacher, K., & Schöberl, M. (2013). A jet space approach to check Pfaffian systems for flatness. In 52nd IEEE conference on decision and control (pp. 2576–2581). https://doi.org/10.1109/CDC.2013.6760270
  • Schöberl, M., Rieger, K., & Schlacher, K. (2010). System parametrization using affine derivative systems. In Proceedings 19th international symposium on mathematical theory of networks and systems (MTNS) (pp. 1737–1743).
  • Schöberl, M., & Schlacher, K. (2010). On parametrizations for a special class of nonlinear systems. IFAC Proceedings Volumes, 43(14), 1261–1266. 8th IFAC symposium on nonlinear control systems. https://doi.org/10.3182/20100901-3-IT-2016.00051
  • Schöberl, M., & Schlacher, K. (2011). On calculating flat outputs for Pfaffian systems by a reduction procedure – demonstrated by means of the VTOL example. In 2011 9th IEEE international conference on control and automation (ICCA) (pp. 477–482). https://doi.org/10.1109/ICCA.2011.6137922
  • Schöberl, M., & Schlacher, K. (2014). On an implicit triangular decomposition of nonlinear control systems that are 1-flat – a constructive approach. Automatica, 50(6), 1649–1655. https://doi.org/10.1016/j.automatica.2014.04.007
  • Schöberl, M., & Schlacher, K. (2015). On geometric properties of triangularizations for nonlinear control systems. In M. K. Camlibel, A. A. Julius, R. Pasumarthy, and J. M. Scherpen (Eds.), Mathematical control theory i (pp. 237–255). Springer International Publishing.

Appendices

Appendix 1. Supplements

A.1. Proof of Theorem 3.3

In Gstöttner et al. (Citation2020a), the following result on the linearisation of (x,u)-flat two-input systems has been shown (corresponding to Corollary 4 therein).

Lemma A.1

If the two-input system (Equation1) possesses an (x,u)-flat output with a certain difference d, then it can be rendered static feedback linearisable by d-fold prolonging a suitably chosen (new) input after a suitable input transformation u¯=Φu(x,u) has been applied.

To prove Theorem 3.3, i.e. to show that two-input systems with d2 can be rendered static feedback linearisable by d-fold prolonging a suitably chosen (new) input, we only have to show that d2 implies (x,u)-flatness. To be precise, we have to show that minimal flat outputs with d2 are (x,u)-flat outputs. Below we will show this by contradiction, i.e. we will show that a flat output with a difference of d2 which explicitly depends on derivatives of the inputs is never a minimal flat output. For that we will utilise a certain relation between the state dimension n of the system, the difference d of a flat output and the ‘generalised’ relative degrees of its components, which we derive in the following.

Consider a flat output y and recall that rj denotes the order of the highest derivative rj of yj needed to parameterise x and u by this flat. For a component of a flat output which does not explicitly depend on a derivative of the inputs we define the relative degree kj by (A1) Lfkj1φj=φkj1j(x),Lfkjφj=φkjj(x,u).(A1) The following lemma provides a relation between the state dimension n, rj and kj.

Lemma A.2

For an (x,u)-flat output y=φ(x,u) of a two input system of the form (Equation1) the relations r1=nk2 and r2=nk1 hold.

Proof.

See Equation (8) in Gstöttner et al. (Citation2020a), i.e. n#K=r1k1 and n#K=r2k2 (where #K=k1+k2), from which readily the relations r1=nk2 and r2=nk1 follow.

With these two relations, we immediately obtain a relation between the difference d=#Rn of an (x,u)-flat output and the relative degrees kj of its components.

Corollary A.3

For an (x,u)-flat output y=φ(x,u) of a two-input system of the form (Equation1) the relation d=nk1k2 holds.

The definition of the relative degree via (EquationA1) requires that φj=φj(x) or φj=φj(x,u) and always yields kj0. However, it turns out that Corollary A.3 (and in fact also Lemma A.2) analogously apply if one or both components of the flat output explicitly depend on derivatives of the inputs, i.e. yj=φj(x,u,u1,,uqj) with φj explicitly depending on uqj1 or uqj2, by setting kj=qj. In other words, Corollary A.3 analogously applies if we interpret an explicit dependence of yj=φj(x,u,u1,,uqj) on the qjth derivative of an input as a negative relative degree of kj=qj. This can be shown as follows. Let y be a flat output of the system (Equation1) and consider the prolonged system x˙=f(x,u)u˙=u1u˙1=u2u˙p1=up,with the state x~=(x,u,u1,,up1), dim(x~)=n~=n+2p and the input u~=up, obtained by p-fold prolonging both inputs of the system. The flat parameterisation x~=Fx~(y[R~1]), u~=Fu~(y[R~]) of the prolonged system with respect to y is easily obtained from the corresponding flat parameterisation of the original system by successive differentiation (we thus have R~=R+p) and it immediately follows that we have d~=#R~n~=#Rn=d, i.e. such total prolongations preserve the difference of every flat output.

Let us consider the case that one component of the flat output explicitly depends on derivatives of the inputs. Without loss of generality (swap the components of the flat output if necessary), we can assume that y1=φ1(x,u) has a relative degree of k10 (we have k1=0 if φ1 explicitly depends on u1 or u2) and y2=φ2(x,u,u1,,uq2) (with φ2 explicitly depending on uq21 or uq22). By q2-fold prolonging both inputs of the system, we obtain the system x˙=f(x,u)u˙=u1u˙1=u2u˙q21=uq2,with the state x~=(x,u,u1,,uq21), dim(x~)=n~=n+2q2 and the input u~=uq2. For this prolonged system, the components of the flat output y=(φ1(x,u),φ2(x,u,u1,,uq2)) only depend on the state and the inputs and thus, Corollary A.3 directly applies. By construction we have k~2=0 since φ2 explicitly depends on u~. The k1th derivative of φ1 explicitly depends on u and therefore we have to differentiate it another q2 times until it explicitly depends on u~ and thus k~1=k1+q2. According to Corollary A.3 we thus have d~=n~k~1k~2=n+2q2(k1+q2)0=nk1+q2. Above we noticed that d=d~, i.e. total prolongations preserve the difference and thus, d=nk1+q2, i.e. exactly the same as Corollary A.3 would yield when we set k2=q2.

The case that both components of the flat output explicitly depend on derivatives of the inputs, i.e. yj=φj(x,u,u1,,uqj) (with φj explicitly depending on uqj1 or uqj2), can be handled analogously and yields d=n+q1+q2, i.e. exactly the same as Corollary A.3 would yield when we set k1=q1 and k2=q2.

With these preliminary results at hand we can now prove that minimal flat outputs with d2 of a two-input system of the form (Equation1) are actually (x,u)-flat outputs. Let y=(φ1(x,u,u1,,uq1),φ2(x,u,u1,,uq2)) be a minimal flat output of (Equation1) with d2. In the following, we show that q1,q20 by contradiction. Without loss of generality, we can assume that q1q2 (swap the components of the flat output if necessary). Assume that q21. According to the above discussed generalisation of Corollary A.3, we have d=n+q1+q2 and thus n+q1+q22. Because of n+q1+q22, we have q12nq2. We can assume n3, since for n = 2, the system would be static feedback linearisable with the state of the system forming a linearising output due to the assumption rank(uf)=2, which is in contradiction with the minimality of y, and for n = 1 the rank condition rank(uf)=2 could not hold. Thus n3 and in turn the first component of the flat output has a (positive) relative degree of k1=q1n+q22. This enables us to replace k1 states of the system by φ1 and its first k11 derivatives by applying the state transformation (A2) x¯1i1=Lfi11φ1(x), i1=1,,k1x¯2i2=ψi2(x), i2=1,,nk1,(A2) where ψi2 are arbitrary functions of the state, chosen such that (EquationA2) is a regular state transformation. By additionally applying the input transformation u¯1=Lfk1φ1, u¯2=g(x,u), with g chosen such that this transformation is invertible with respect to u, we obtain (A3) x¯˙11=x¯12x¯˙12=x¯13x¯˙1k1=u¯1x¯˙2i2=f¯2i2(x¯1,x¯2,u¯1,u¯2),i2=1,,nk1.(A3) Recall that we have k1n+q22 and thus dim(x¯2)=nk12q2. Due to the assumption q21, we thus have dim(x¯2)1. However, for dim(x¯2)=1, (EquationA3) reads x¯˙11=x¯12x¯˙12=x¯13x¯˙1n1=u¯1x¯˙21=f¯21(x¯1,x¯2,u¯1,u¯2),so the system would be static feedback linearisable with (x¯11,x¯21) forming a linearising output, which is in contradiction with the minimality of the flat output y. For dim(x¯2)=0, the system would consist of a single integrator chain of length n, which actually contradicts with rank(uf)=2. We thus have q20 and because of q1q2, also q10. In conclusion, every minimal flat output with a difference of d2 is an (x,u)-flat output. Lemma A.1 therefore guarantees that a system with d2 can be rendered static feedback linearisable by d-fold prolonging a suitable chosen (new) input after a suitable input transformation u¯=Φu(x,u) has been applied, which completes the proof.

A.2. Proof of Lemma 5.1

By assumption, we have [f,Dk12]Dk11 and Dk1=Dk11+[f,Dk11] with Dk12 and Dk11 being involutive. For every vector field wDk12 and vDk11 the Jacobi identity [v,[w,f]Dk11]Dk11Dk1+[f,[v,w]Dk11]Dk1+[w,[f,v]Dk]=0holds, from which it follows that Dk12C(Dk1). We furthermore have Dk122Dk112Dk1. For any vector fields v1,v2 such that Dk11=Dk12+span{v1,v2} we thus have Dk1=Dk11+span{[v1,f],[v2,f]}. By assumption we have Dk11C(Dk1), which in turn implies that at least one of the vector fields [v1,[v1,f]], [v1,[v2,f]] or [v2,[v2,f]]Footnote8 in (Equation5), i.e.  (A4) (α1)2[v1,[v1,f]]+2α1α2[v1,[v2,f]]+(α2)2[v2,[v2,f]]!Dk1,(A4) is not contained in Dk1. If they are all linearly independent modulo Dk1, i.e. dim(Dk1+[Dk11,Dk1])=dim(Dk1)+3, then in order for (EquationA4) to hold, each coefficient has to be zero, i.e. (α1)2=0, 2α1α2=0 and (α2)2=0, which only admits the trivial solution α1=α2=0. So we only have to consider the cases dim(Dk1+[Dk11,Dk1])=dim(Dk1)+2 and dim(Dk1+[Dk11,Dk1])=dim(Dk1)+1, i.e. those cases in which [Dk11,Dk1] yields 2 new directions or 1 new direction with respect to Dk1, respectively.

Case 1: Let us first consider the case dim(Dk1+[Dk11,Dk1])=dim(Dk1)+2, i.e. H=Dk1+[Dk11,Dk1]=Dk1+span{w1,w2}, where w1 and w2 are a suitable selection of the three vector fields [v1,[v1,f]], [v1,[v2,f]] and [v2,[v2,f]]. We distinguish between the following subcases:

  1. H=Dk1+span{[v1,[v1,f]],[v1,[v2,f]]} and thus [v2,[v2,f]]=κ1[v1,[v1,f]]+κ2[v1,[v2,f]] mod Dk1, which inserted into (EquationA4) yields ((α1)2+(α2)2κ1)[v1,[v1,f]]+(2α1α2+(α2)2κ2)[v1,[v2,f]]!Dk1. Since [v1,[v1,f]] and [v1,[v2,f]] are by assumption not collinear modulo D1, the (up to a multiplicative factor) only non-trivial solution is α1=κ2 and α2=2, provided that (κ2)2+4κ1=0, otherwise, no non-trivial solution exists. If subcase (1) is not applicable, then the vector fields [v1,[v1,f]] and [v1,[v2,f]] are collinear modulo Dk1, i.e. [v1,[v2,f]]=κ[v1,[v1,f]] mod Dk1 or [v1,[v1,f]]Dk1.

  2. The subcase [v1,[v2,f]]=κ[v1,[v1,f]] mod Dk1 yields ((α1)2+2α1α2κ)[v1,[v1,f]]+(α2)2[v2,[v2,f]]!Dk1.Since the vector fields [v1,[v1,f]] and [v2,[v2,f]] are by assumption not collinear modulo Dk1, the condition only admits the trivial solution α1=α2=0.

  3. The remaining subcase [v1,[v1,f]]Dk1 yields 2α1α2[v1,[v2,f]]+(α2)2[v2,[v2,f]]!Dk1.Since [v1,[v2,f]] and [v2,[v2,f]] are by assumption not collinear mod Dk1, the (up to a multiplicative factor) only non-trivial solution is α1=1 and α2=0.

Case 2: Consider the case dim(Dk1+[Dk11,Dk1])=dim(Dk1)+1, i.e. H=Dk1+[Dk11,Dk1]=Dk1+span{w1}, where w1 is either [v1,[v1,f]], [v1,[v2,f]] or [v2,[v2,f]]. We distinguish between the following subcases:

  1. H=Dk1+span{[v1,[v1,f]]} and thus [v1,[v2,f]]=κ1[v1,[v1,f]] mod Dk1 and [v2,[v2,f]]=κ2[v1,[v1,f]] mod Dk1, which inserted into (EquationA4) yields ((α1)2+2α1α2κ1+(α2)2κ2)[v1,[v1,f]]!Dk1.The (up to a multiplicative factor) only non-trivial solutions are α1=κ1±(κ1)2κ2 and α2=1. If subcase (1) is not applicable, i.e. [v1,[v1,f]]Dk1, we either have [v2,[v2,f]]=κ[v1,[v2,f]] mod Dk1, or [v1,[v1,f]],[v1,[v2,f]]Dk1.

  2. The subcase [v1,[v1,f]]Dk1 and [v2,[v2,f]]=κ[v1,[v2,f]] mod Dk1 yields (2α1α2+(α2)2κ)[v1,[v2,f]]!Dk1.The (up to a multiplicative factor) only non-trivial solutions are α1=1, α2=0 and α1=κ, α2=2.

  3. The remaining subcase [v1,[v1,f]],[v1,[v2,f]]Dk1 yields (α2)2[v2,[v2,f]]!Dk1,and the (up to a multiplicative factor) only non-trivial solution is α1=1, α2=0.

In conclusion, if dim(Dk1+[Dk11,Dk1])=dim(Dk1)+3, there exists no non-trivial solution. If dim(Dk1+[Dk11,Dk1])=dim(Dk1)+2, there either exists (up to a multiplicative factor) only one solution or no non-trivial solution. If dim(Dk1+[Dk11,Dk1])=dim(Dk1)+1, there exist (up to a multiplicative factor) at most two solutions or no non-trivial solutions (complex solutions are not relevant).

A.3. The special case k1=1 with Δi of the form (17)

By assumption we have k1=1 and adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}, i.e. the distribution D1 is non-involutive and the involutive distributions Δi are of the form (Equation17). It follows that we necessarily have D0C(D1). Indeed, D0C(D1) would imply [u¯1,[f¯,u¯2]]D1 and in turn either Δ2T(X×U)=D1, which contradicts with D1 being non-involutive, or Δ2=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]} and then, due to adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}, Δ3T(X×U)=D1 would follow, which again contradicts with D1 being non-involutive. Thus [u¯1,[f¯,u¯2]]D1 and therefore D0C(D1).

Thus we have to show that the system necessarily meets item 2b, i.e. that there exists a non-vanishing vector field vcD0 such that vcC(E1) where E1 is defined as E1=D0+span{[vc,f¯]}. Let us show that the vector field vc=u¯2 meets this criterion. The vector field vc=u¯2 meets u¯2D0 and yields E1=D0+span{[f¯,u¯2]}=span{u¯1,u¯2,[f¯,u¯2]}. The involutivity of Δ1 implies that u¯2C(E1). Therefore, vc=u¯2 and E1=span{u¯1,u¯2,[f¯,u¯2]}.

Above, we have deduced that [u¯1,[f¯,u¯2]]D1, which because of E1D1 implies that [u¯1,[f¯,u¯2]]E1 and thus E1 is non-involutive. So we have to show next that item 3a is necessarily met.

We have E1Δ2T(X×U) (see (Equation17)), from which the involutive closure of E1 follows as E¯1=Δ2T(X×U)=span{u¯1,u¯2,[f¯,u¯2],[u¯1,[f¯,u¯2]]} and thus 3a.I is met. Note that we necessarily have E¯1T(X×U). Indeed, since dim(E1)=3 and dim(E¯1)=4, E¯1=T(X×U) would imply n = 2 and thus D1=T(X×U), which is in contradiction with D1 being non-involutive. Therefore, we necessarily have E¯1T(X×U), based on which we next show that necessarily 3a.II is met, i.e. that necessarily dim([f¯,E1]+E¯1)=dim(E¯1)+1. Since adf¯2u¯2Δ2T(X×U)=E¯1, the condition holds if and only if [f¯,u¯1]E¯1, which can be shown by contradiction. Assume that [f¯,u¯1]E¯1 and thus [f¯,E1]+E¯1=E¯1. Based on the Jacobi identity, it can then be shown that this would imply [f¯,E¯1]E¯1. However, since E¯1T(X×U), this would in turn imply that the system contains an autonomous subsystem, which is in contradiction with the assumption that the system is flat.

With F2=E¯1=Δ2T(X×U), it follows that the distributions Fi=Fi1+[f¯,Fi1], i3 constructed in item 5 and the distributions Δi are related via Δi=span{u¯21,u¯11}+Fi and thus F3=Δ3T(X×U)Fs=ΔsT(X×U)=T(X×U).So these distributions are indeed involutive and there indeed exists an integer s such that Es=T(X×U), which completes the necessity part of the proof for the special case k1=1 with adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}.

Appendix 2. Proofs of Lemmas

A.4. Proof of Lemma 7.1

We have [fp,u¯j]=u¯jf¯i(x,u¯)xi=[f¯,u¯j] with f¯=f¯i(x,u¯)xi and thus Δ1=span{u¯21,u¯11,u¯2,[f¯,u¯2]} and in turn Δ2=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2],[fp,[f¯,u¯2]]}.We have [fp,[f¯,u¯2]]=adf¯2u¯2+u¯11[u¯1,[f¯,u¯2]]+u¯21[u¯11,[f¯,u¯2]]=adf¯2u¯2+u¯11[u¯1,[f¯,u¯2]] mod Δ1. The involutivity of Δ2 implies that adf¯2u¯2 and [u¯1,[f¯,u¯2]] are collinear modulo span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}. In case of k12, we certainly have adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}. Indeed, assume that k12 and adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}. We then have Δ2=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2],[u¯1,[f¯,u¯2]]D1}(where [u¯1,[f¯,u¯2]]D1 due to the involutivity of D1). Thus either Δ2=span{u¯21,u¯11}+D1 or Δ2=span{u¯21,u¯11}+span{u¯1,u¯2,[f¯,u¯2]}. It follows that the first case would lead to Δi=span{u¯21,u¯11}+Di1 which for i=k1+1 would imply that Dk1 is involutive, contradicting with Dk1 being non-involutive. The second case, i.e. Δ2=span{u¯21,u¯11}+span{u¯1,u¯2,[f¯,u¯2]} would lead to Δ3=span{u¯21,u¯11}+D1 and in turn Δi=span{u¯21,u¯11}+Di2 which for i=k1+2 would imply that Dk1 is involutive, again contradicting with Dk1 being non-involutive. Thus, in case of k12 we always have adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]}, and thus Δ2=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2],adf¯2u¯2}.The form (Equation16) then easily follows from the involutivity of the distributions Δ2,Δ3, and the fact that fp and f¯ coincide modulo span{u¯21,u¯11}Δ2,Δ3,

However, for k1=1, it can indeed happen that adf¯2u¯2span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2]} and thus Δ2=span{u¯21,u¯11,u¯1,u¯2,[f¯,u¯2],[u¯1,[f¯,u¯2]]}(in which case we necessarily have [u¯1,[f¯,u¯2]]D1). In this case, due to the involutivity of the distributions Δ2,Δ3, and the fact that fp and f¯ coincide modulo span{u¯21,u¯11}Δ2,Δ3, the form (Equation17) follows.

A.5. Proof of Lemma 7.2

Under the assumption d = 2 with Dk11C(Dk1), the distribution Ik1=Δk1+1Dk1 is involutive, which can be shown by contradiction. Assume that Ik1=Δk1+1Dk1 is non-involutive. From (Equation16), it follows that we haveFootnote9 Ik1=Δk1+1Dk1=Dk11+span{adf¯k1u¯2}.Since Ik1Δk1+1T(X×U) and Δk1+1T(X×U)=Dk11+span{adf¯k1u¯2,adf¯k1+1u¯2}is involutive, we then have I¯k1=Δk1+1T(X×U), i.e. Ik11I¯k1. We have Ik11Dk1 and since Dk11C(Dk1) we have [Dk11,Ik1]Dk1. However, since Ik1 is assumed to be non-involutive we necessarily have [Dk11,Ik1]Ik1 and thus, the direction which completes Ik1 to its involutive closure is contained in Dk1, which would lead to I¯k1=Dk1 and contradict with Dk1 being non-involutive. Thus Ik1 is indeed involutive.

From the Jacobi identity [adf¯k1u¯2,[adf¯k11u¯1,f¯]]Dk1(1)+[f¯,[adf¯k1u¯2,adf¯k11u¯1]Ik1]Dk1+span{adf¯k1+1u¯2}+[adf¯k11u¯1,[f¯,adf¯k1u¯2]]Δk1+1T(X×U)Dk1+span{adf¯k1+1u¯2}=0,where we have used the involutivity of Ik1 for the second term, it follows that the vector field [adf¯k1u¯1,adf¯k1u¯2], which completes Dk1 to Dk1(1), is contained in Dk1+span{adf¯k1+1u¯2}, and thus Dk1(1)=Dk1+span{adf¯k1+1u¯2}, as desired.

A.6. Proof of Lemma 7.3

The vector field adf¯k21u¯2Ek21 is a Cauchy characteristic vector field of Ek2. Indeed, we have (see (Equation19) and (Equation20)) (A5) Ek2=Ek21+span{adf¯k2u¯2}Δk2T(X×U)+span{adf¯k21u¯1}.(A5) From the Jacobi identity [adf¯k21u¯2,[f¯,adf¯k22u¯1]]+[adf¯k22u¯1,[adf¯k21u¯2,f¯]]Δk2T(X×U)Ek2+[f¯,[adf¯k22u¯1,adf¯k21u¯2]Ek21]Ek2=0,it follows that [adf¯k21u¯2,adf¯k21u¯1]Ek2, where we have used the fact that [f¯,Ek21]Ek2 and that Δk2T(X×U) and Ek21 are involutive. The involutivity of Δk2T(X×U) together with [adf¯k21u¯2,adf¯k21u¯1]Ek2 implies that adf¯k21u¯2C(Ek2).

The distribution Ek2 is by assumption non-involutive. Since Ek2Δk2+1T(X×U) and Δk2+1T(X×U) is involutive it follows that E¯k2=Δk2+1T(X×U)=Ek2+span{adf¯k2+1u¯2}.

In case of k2k1+2 it immediately follows from the Jacobi identity and the way the distributions Ei are constructed that Ek22C(Ek2). We have Ek21=Ek22+span{adf¯k22u¯1,adf¯k21u¯2} and we have just shown that adf¯k21u¯2C(Ek2). Since by assumption Ek21C(Ek2) it follows that adf¯k22u¯1C(Ek2). Since adf¯k22u¯1 is contained in the involutive subdistribution Δk2T(X×U)1Ek2 (see (A5)), it follows that the vector field [adf¯k22u¯1,adf¯k21u¯1] cannot be contained in Ek2 and thus E¯k2=Ek2+span{[adf¯k22u¯1,adf¯k21u¯1]}, which in turn implies that the vector fields [adf¯k2u¯2,adf¯k21u¯1] and [adf¯k22u¯1,adf¯k21u¯1] are collinear modulo Ek2, i.e. [adf¯k2u¯2,adf¯k21u¯1]=λ[adf¯k22u¯1,adf¯k21u¯1] mod Ek2. From the latter it follows that [adf¯k2u¯2λadf¯k22u¯1,adf¯k21u¯1]Ek2, which because of adf¯k2u¯2λadf¯k22u¯1Δk2T(X×U) implies that adf¯k2u¯2λadf¯k22u¯1C(Ek2). It follows that C(Ek2)=Ek22+span{adf¯k21u¯2,adf¯k2u¯2λadf¯k22u¯1}2Ek2, i.e. further Cauchy characteristic vector fields which are linearly independent of those we have already found cannot exist, otherwise Ek2 would be involutive.

In case of k2=k1+1, the distribution Ek22 does not exist. However, we have the involutive distribution Δk1Dk11=Dk12+span{adf¯k11u¯2} and we have Ek1=Dk11+span{adf¯k1u¯2} in this case. Thus it follows that [f¯,Δk1Dk11]Ek1. Based on the latter by applying the Jacobi identity, it can be shown that Δk1Dk11C(Ek2). Repeating the proof above with Ek22 replaced by Dk11Δk1 gives the desired result that in case of k2=k1+1 we have C(Ek2)=Dk11Δk1+span{adf¯k21u¯2,adf¯k2u¯2λadf¯k22u¯1}.

A.7. Proof of Lemma 7.4

By assumption we have Gk1C(Gk), which actually implies Gk1=C(Gk). Indeed, since Gk is non-involutive, we necessarily have dim(C(Gk))dim(Gk)2. Because of Gk1C(Gk) and Gk12Gk, we thus have Gk1=C(Gk). Apply a change of coordinates (z3,z2,z1)=Φ(x,u) such that Gk1 and G¯k get straightened out simultaneously, i.e. such that in the new coordinates we have Gk1=span{z1},G¯k=span{z1,z2}.6ptSince Gk1=C(Gk), in these coordinates there exists a basis for Gk of the form Gk=Gk1+span{v1,v2} with the two vector fields vj being of the form vj=vji2(z3,z2)z2i2, i2=1,,dim(z2) (where dim(z2)=3).

In case of G¯k=T(X×U) (we have dim(z3)=0 and thus simply vj=vji2(z2)z2i2 in this case) choose any function ψ(z2), dψ0 and define the vector field vc=(dψv2)v1(dψv1)v2, which is nonzero due to the non-involutivity of Gk (indeed, dψv1=dψv20 would yield Gk=span{dψ} which contradicts with Gk being non-involutive). The distribution Hk=Gk1+span{vc} is obviously involutive and we by construction have Gk11Hk1Gk. Note that there exist infinitely many valid choices for the function ψ and different choices yield in general different distributions Hk.Footnote10

Next, let us consider the case G¯kT(X×U). Note that because of [f,Gk1]GkG¯k, in the above introduced coordinates the vector field f is of the form f=f3i3(z3,z2)z3i3+f2i2(z3,z2,z1)z2i2+f1i1(z3,z2,z1)z1i1.We thus have [f,Gk]+G¯k=G¯k+span{[v1,f3],[v2,f3]},where f3=f3i3(z3,z2)z3i3. Since by assumption dim([f,Gk]+G¯k)=dim(G¯k)+1, the vector fields [v1,f3] and [v2,f3] are collinear modulo G¯k. Without loss of generality, we can assume that [v1,f3]G¯k, implying that there exists a function α such that [v2,f3]=α[v1,f3] mod G¯k. This function only depends on z3 and z2 since the vector fields v1,v2 and f3 only depend on z3 and z2, i.e. α=α(z3,z2). Define the distribution Hk=Gk1+span{v2αv1}. It is immediate that this distribution is involutive and since v1 and v2 are independent we have Gk11Hk1Gk.Footnote11

So in any of the two cases, i.e. G¯k=T(X×U) or G¯kT(X×U), we have shown that there exists an involutive distribution Hk=Gk1+span{v} with some vector field vGk, vGk1. What is left to do is to show that Hk+[f,Hk]=G¯k. By construction, we have [f,Hk]G¯k, so we only have to show that [f,v]Gk, which can be shown by contradiction. Assume that [f,v]Gk. Due to the Jacobi identity, for every vector field wGk1 we then have [w,[f,v]Gk]Gk+[v,[w,f]Gk]+[f,[v,w]Hk]Gk=0,(where we have used that Gk1C(Gk) and that [f,v]Gk implies [f,Hk]Gk, and the involutivity of Hk). However, this implies [v,[w,f]]Gk for every wGk1, which in turn would imply that Gk is involutive and contradict with Gk being non-involutive.

A.8. Proof of Lemma 7.5

We have Dk1=Dk11+span{v1,v2} and since Dk11C(Dk1) we have Dk1(1)=Dk11+span{v1,v2,[v1,v2]}. Based on the Jacobi identity it can be shown that Dk11C(Dk1(1)). Since by assumption dim(D¯k1)=dim(Dk1)+2, the vector fields [v1,[v1,v2]] and [v2,[v1,v2]] are collinear modulo Dk1(1). Without loss of generality, we can assume that [v1,[v1,v2]]Dk1(1) and thus [v2,[v1,v2]]=α[v1,[v1,v2]] mod Dk1(1). The vector field v=v2αv1 by construction meets vDk1, vDk11 and [v,[v1,v2]]=0 mod Dk1(1), implying that vC(Dk1(1)). It follows that C(Dk1(1))=Dk11+span{v} (the existence of further Cauchy characteristic vector fields would contradict with Dk1(1) being non-involutive) and thus Dk111C(Dk1(1))1Dk1.

That C(Dk1(1))+[f,C(Dk1(1))]=Dk1(1) can be shown as follows. We have Dk11Dk1(1) and by assumption [f,C(Dk1(1))]Dk1(1), and we just have shown that C(Dk1(1))=Dk11+span{v} with vDk1, vDk11. Therefore, C(Dk1(1))+[f,C(Dk1(1))]=Dk1(1) holds if and only if [f,v]Dk1, which can be shown by contradiction. Assume that [f,v]Dk1. Due to the Jacobi identity, for every vector field wDk11 we then have [w,[f,v]Dk1]Dk1+[v,[w,f]Dk1]+[f,[v,w]C(Dk1(1))]Dk1=0,(where we have used that Dk11C(Dk1) and that [f,v]Dk1 implies [f,C(Dk1(1))]Dk1, and the involutivity of C(Dk1(1))). However, this implies that [v,[w,f]]Dk1 for every wDk11, which in turn would imply that Dk1 is involutive and contradict with Dk1 being non-involutive.